114
Research Collection Doctoral Thesis Structure, hydration, and lubricity of surface-attached dextran Author(s): Goren, Tolga Publication Date: 2014 Permanent Link: https://doi.org/10.3929/ethz-a-010105231 Rights / License: In Copyright - Non-Commercial Use Permitted This page was generated automatically upon download from the ETH Zurich Research Collection . For more information please consult the Terms of use . ETH Library

In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

Embed Size (px)

Citation preview

Page 1: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

Research Collection

Doctoral Thesis

Structure, hydration, and lubricity of surface-attached dextran

Author(s): Goren, Tolga

Publication Date: 2014

Permanent Link: https://doi.org/10.3929/ethz-a-010105231

Rights / License: In Copyright - Non-Commercial Use Permitted

This page was generated automatically upon download from the ETH Zurich Research Collection. For moreinformation please consult the Terms of use.

ETH Library

Page 2: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

DISS ETH NO. 21777

Structure, Hydration, and Lubricity of Surface-

Attached Dextran

A dissertation submitted to the

ETH Zürich

for the degree of

Doctor of Sciences

presented by

Tolga Goren

Master of Science, Rensselaer Polytechnic Institute

born 21 May 1987

accepted on the recommendation of

Prof. Dr. Nicholas D. Spencer, examiner

Prof. Dr. Mark Rutland, co-examiner

Dr. Rowena Crockett, co-examiner

2014

Page 3: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

2

Contents

Contents ....................................................................................................................................................... 2

Abstract ....................................................................................................................................................... 5

Zusammenfassung ....................................................................................................................................... 6

Acknowledgments ....................................................................................................................................... 8

Chapter 1: Introduction ............................................................................................................................ 9

1.1 Aqueous lubrication in nature .......................................................................................................... 9

1.2 Polysaccharides in aqueous solutions ............................................................................................ 10

1.3 Polymer brushes ............................................................................................................................ 10

1.4 Scope of the thesis ......................................................................................................................... 13

1.5 References ..................................................................................................................................... 13

Chapter 2: Experimental Methods ......................................................................................................... 18

2.1 Optical characterization of thin films ............................................................................................. 18

2.1.1 Principles and models ............................................................................................................ 18

2.1.2 Variable-angle spectroscopic ellipsometry (VASE)............................................................... 19

2.1.3 Optical waveguide lightmode spectroscopy (OWLS) ............................................................ 20

2.1.4 Transmission interference adsorption sensing (TInAS) ......................................................... 20

2.2 Measurement of friction with the atomic force microscope ........................................................... 21

2.3 Remarks on computer simulations ................................................................................................. 23

2.4 References ..................................................................................................................................... 24

Chapter 3: Load-induced transitions in the lubricity of adsorbed poly(L-lysine)-g-dextran as a

function of polysaccharide chain density ............................................................................................... 26

3.1 Abstract.......................................................................................................................................... 26

3.2 Introduction ................................................................................................................................... 27

3.3 Materials and Methods................................................................................................................... 29

3.3.1 Poly(L-lysine)-graft-dextran (PLL-g-dex) ............................................................................. 29

3.3.2 Gradient preparation .............................................................................................................. 30

3.3.3 Atomic force microscope measurements ............................................................................... 32

3.4 Results ........................................................................................................................................... 34

3.4.1 Measurement of L/2Rg ........................................................................................................... 34

3.4.2 Friction measurements ........................................................................................................... 35

3.5 Discussion ...................................................................................................................................... 38

Page 4: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

3

3.5.1 Effect of L/2Rg ....................................................................................................................... 38

3.5.2 Effect of load ......................................................................................................................... 39

3.6 Conclusions ................................................................................................................................... 40

3.7 Supporting data: calibration of lateral-force measurements ........................................................... 41

3.8 References ..................................................................................................................................... 44

Chapter 4: Influence of solutes on hydration and lubricity of dextran brushes ................................. 51

4.1 Abstract.......................................................................................................................................... 51

4.2 Introduction ................................................................................................................................... 51

4.3 Experimental .................................................................................................................................. 53

4.3.1 PLL-g-dextran brushes ........................................................................................................... 53

4.3.2 Optical measurements of brush thickness and refractive index .............................................. 53

4.3.3 Colloidal-probe lateral force measurements ........................................................................... 55

4.4 Results and discussion ................................................................................................................... 55

4.4.1 Brush solvation ...................................................................................................................... 55

4.4.2 Aqueous lubricity ................................................................................................................... 56

4.4.3 Influence of grafting density .................................................................................................. 56

4.4.3 Influence of solutes ................................................................................................................ 57

4.5 Conclusions ................................................................................................................................... 59

4.6 References ..................................................................................................................................... 60

Chapter 5: Impact of chain morphology on the lubricity of surface-grafted polysaccharides .......... 64

5.1 Abstract.......................................................................................................................................... 64

5.2 Introduction ................................................................................................................................... 65

5.3 Materials and methods ................................................................................................................... 67

5.3.1 Synthesis of poly(allylamine)-graft-perfluorophenylazide (PAAm-g-PFPA) ........................ 67

5.3.2 Substrate preparation and dextran attachment ........................................................................ 67

5.3.3 Characterization by variable-angle spectroscopic ellipsometry (VASE) ............................... 68

5.3.4 Atomic force microscopy measurements ............................................................................... 69

5.4 Results and discussion ................................................................................................................... 70

5.4.1 Gradient fabrication ............................................................................................................... 70

5.4.2 Friction measurements ........................................................................................................... 74

5.5 Conclusions ................................................................................................................................... 79

5.6 Supplementary information ........................................................................................................... 79

5.7 References ..................................................................................................................................... 81

Chapter 6: Hydration and conformation in dextran and PEG brushes .............................................. 87

Page 5: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

4

6.1 Abstract.......................................................................................................................................... 87

6.2 Introduction ................................................................................................................................... 87

6.3 Materials and methods ................................................................................................................... 89

6.3.1 Poly(L-lysine)-graft-dextran (PLL-g-dextran) and poly(L-lysine)-graft-poly(ethylene

glycol) (PLL-g-PEG) ............................................................................................................................. 89

6.3.2 Transmission interference adsorption sensing (TInAS) ......................................................... 90

6.3.3 Coarse-grained dextran simulations ....................................................................................... 90

6.3.4 ATR crystal preparation and brush attachment ...................................................................... 91

6.4 Results and discussion ................................................................................................................... 92

6.4.1 TInAS thickness and refractive index measurements of PLL-g-dex ...................................... 92

6.4.2 Simulations ............................................................................................................................ 94

6.4.3 ATR-IR .................................................................................................................................. 98

6.5 Conclusions ................................................................................................................................. 101

6.6 Supporting information ................................................................................................................ 102

6.7 References ................................................................................................................................... 104

Chapter 7: Thesis Summary and Outlook ........................................................................................... 110

Page 6: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

5

Abstract

Tribosystems found in nature consist of biomolecule-coated surfaces interacting under

aqueous conditions, often operating in the boundary-lubrication regime. Highly glycosylated

biomolecules play a key role in lending these tribosystems their characteristically low friction

and wear rates. While synthetic biomimetic coatings such as polymer brushes have shown

promise in achieving ultralow friction and wear, the specific role of sugars in the friction of

natural glycoproteins and synthetic polysaccharide brushes is not yet understood. This work

aims to determine the influence of structure and conformation on the hydration and lubricity

of dextran-based brushes and brush-like films. Colloidal-probe friction-force measurements

of dextran brush-density gradients showed friction coefficients which varied over several

orders of magnitude with applied load as well as chain density, as inter-chain hydrogen

bonding lent the polysaccharide brush resistance against collapse, yet exacerbated friction

forces upon collapse or interpenetration with a brush on the opposing surface. The addition of

kosmotropic solutes was shown to reduce the load-bearing capacity of dextran brushes,

disrupting the inter-chain hydrogen bonding and moving the transition to the high-friction

regime to lower applied loads. An alternative grafting system was used to produce

disordered, brush-like dextran films with higher chain densities, which showed less extreme

friction behavior and an inverted influence of chain density on friction at high loads

compared with dextran brushes, highlighting the importance of the ordered chain

conformation in both resisting collapse and sticking upon collapse. The hydration of dextran

brushes and the conformation dependence on chain density was also investigated, showing

key differences in chain stiffness and equilibrium height compared to conventional polymer

brushes. These results cumulatively enhance our understanding of potential mechanisms in

biolubrication, showing the importance of control of film structure and environmental

parameters, such as loading and salt concentration, in achieving the desired friction regime.

Page 7: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

6

Zusammenfassung

Viele natürliche Reibungssysteme bestehen aus mit Biomolekülen beschichteten

Oberflächen, welche in wässriger Umgebung im Bereich der sogenannten

Grenzflächenschmierung miteinander wechselwirken. Dabei spielen stark glykosylierte

Biomoleküle eine Schlüsselrolle. Sie verleihen diesen Reibungssystemen ihre charakteristisch

tiefe Reibung und Verschleissraten. Während sich gezeigt hat, dass mit synthetischen

biomimetischen Beschichtungen wie Polymerbürsten ultraniedrige Reibung und

ultraniedriger Verschleiss erzielt werden kann, wurde die spezifische Rolle von Zuckern in

der Reibung, z.B. in natürlichen Glykoproteinen und synthetischen Polysaccharid-Bürsten

noch nicht verstanden. In dieser Arbeit wurde der Einfluss der Struktur und der Konformation

von Dextran-basierten Bürsten und bürstenähnlichen Filmen auf deren Hydratation und

Schmierfähigkeit untersucht. Mit einer Kolloidspitze durchgeführte Reibungskraftmessungen

auf Dextran-Dichtegradienten haben gezeigt, dass der Reibungskoeffizient abhängig von der

aufgebrachten Kraft und Kettendichte über mehrere Grössenordnungen variiert. Dies wird

beobachtet, weil die Bildung von Wasserstoffbrücken zwischen den Ketten der

Polysaccharid-Bürste einerseits einen gewissen Widerstand gegen das Kollabieren verleiht,

andererseits die Reibungskräfte erhöht, wenn die Bürste kollabiert oder sich die Bürsten von

gegenüberliegenden Oberflächen gegenseitig durchdringen. Es wurde gezeigt, dass der

Zusatz von kosmotropischen gelösten Substanzen dazu führt, dass das Tragvermögen von

Dextranbürsten reduziert wird, da so die Bildung von Wasserstoffbrücken gestört wird, so

dass sich bereits bei tieferen Lasten eine hohe Reibung zeigt. Mit einem alternativen Ansatz

wurden ungeordnete, bürstenähnliche Dextranfilme mit höheren Kettendichten hergestellt,

welche ein weniger extremes Reibungsverhalten zeigten. Auch kehrte sich im Vergleich zu

Dektranbürsten der Einfluss der Kettendichte bei hohen Lasten um, was zeigt wie wichtig die

Ordnung und Konformation der Ketten beim Verhindern des Kollapses, sowie beim Haften

aufgrund eines Kollaps ist. Eine Untersuchung der Hydratation, sowie der Abhängigkeit der

Konformation von der Kettendichte zeigte, dass gegenüber konventionellen Polymerbürsten

entscheidende Unterschiede in der Steifigkeit der Kette und in der Schichtdicke bestehen.

Diese Resultate verbessern unser Verständnis der Mechanismen, welche bei der Schmierung

von Biosystemen wirken stark und sie zeigen, wie wichtig es ist, die Struktur des

Schmierfilms und die Umgebungsbedingungen, wie z.B. die Belastung oder die

Page 8: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

7

Salzkonzentration zu kontrollieren, wenn ein gewünschter Reibungsbereich erreicht werden

soll.

Page 9: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

8

Acknowledgments

For their invaluable support, I would like to acknowledge Prof. Nicholas D. Spencer and Dr.

Rowena Crockett, who gave me the chance to pursue this project and who supervised me

throughout; Dr. Kenneth Rosenberg, who co-supervised the first year of my work and

contributed substantially to the work described in Chapter 3; Dr. Venkataraman

Nagaiyanallur, who contributed substantially to the work described in Chapter 6; and the

other members of the Laboratory of Surface Science and Technology, past and present, for

their support throughout.

Outside our own group, I also thank Prof. Mark Rutland, for flying in to be my co-examiner,

and for providing helpful feedback about the thesis; Prof. Walter Steurer, for chairing my

thesis defense; Dr. Patrick Ilg, for valuable suggestions about the simulations described in

Chapter 6; Prof. Takumi Sannomiya, for assistance with the TInAS-FIT software; Martin

Elsener and Markus Müller, for technical support; Dr. Marc Petitmermet, for helping with the

LSST Wiki creation; Dr. Ben Seeber, for SEM support; Dr. Thomas Schweizer, for assistance

with DSC and TGA measurements; Michael Horisberger of the Paul Scherrer Institute, for

help with substrate coating and TInAS sensor production; Dr. Martin Willeke, for guiding me

in fulfilling my teaching responsibilities; and the students I supervised: Gianluca Etienne,

Manuel Baumgartner, Selmar Binder, Jana Segmehl, Kotryna Bloznelyte, Marius Staub, Kay

Sanvito, and Hendrik Spanke.

I am also grateful for financial support throughout my study from the Swiss National Science

Foundation and from ETH.

Page 10: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

9

Chapter 1: Introduction

1.1 Aqueous lubrication in nature

Look around you! You don’t have to look far: the surfaces of your eyeballs consist of

epithelial cells bearing a layer of mucins, which extend out into a protein-rich tear film.

When you blink, your upper eyelid slides along this surface at about 1 cm/s, applying a

pressure of 1-5 kPa onto the cornea – a tribological process in the boundary-lubrication

regime, which mechanically wipes the surface of the eye while renewing the tear film.1 But

you don’t even have to look that far: blood cells are being transported around your retinas

through vessels and capillaries, which in some cases are narrower than the diameter of a

single red blood cell. Erythrocytes are physically deformed into plug shapes as they squeeze

through the narrowest capillaries, lubricated by the surrounding blood plasma and by the

carbohydrate-rich glycocalyx coating the capillary walls.2

The aqueous lubrication mechanisms found in these and other natural tribosystems consist of

complex, mutually interacting biological structures. The characteristic low-friction and low-

wear behavior of oil-based lubricants operating in the fluid-film and elastrohydrodynamic

lubrication regimes are difficult to achieve with aqueous lubricants alone, due to their

relatively low viscosity and tendency to be squeezed out of the contact area.3 Natural

tribosystems have evolved to meet the challenge by deploying various lubricant additives,

such as hyaluronan,4 and utilizing surfaces of hierarchical composition bearing glycosylated

biomolecules, such as lubricin,5 which both help to trap the water film under pressure and

directly reduce friction upon direct contact under hydrated conditions. The combination

forms a tribosystem that operates in the boundary regime but with friction and wear rates

resembling fluid film or elastrohydrodynamic lubrication. Understanding the physical

phenomena that these systems exploit is useful for industries where aqueous lubrication is a

necessity (including biomedical, civil and aquatic engineering) or a desirable non-toxic,

environmentally friendly alternative to oil-based lubrication (including food and

pharmaceutical engineering). While existing systems fall short of true biomimicry, synthetic

aqueous lubrication solutions have come into use both as model systems for study and in a

variety of applications.

Page 11: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

10

Glycoproteins, which consist of a protein bearing multiple saccharide chains, are an

important class of natural lubricant additives.6 The hydrophilic sugars serve to trap large

amounts of water under pressure, and elaborate glycosylation mechanisms produce glycans of

controlled, heterogeneous composition, which gives fine control over the length, stiffness,

and chemical activity of the chains, and thereby their role in controlling aqueous lubrication.7

1.2 Polysaccharides in aqueous solutions

The behavior of polysaccharides in water is closely interrelated with their hydration and their

inter-chain interactions. Glucose is extremely water-soluble, chemically resembling a water

cluster, aside from small hydrophobic patches on the two carbohydrate ring faces. Despite

this, many glucans form insoluble structures by virtue of rigid glycosidic linkages and

coordinated inter-chain hydrogen bonding, which can cooperate to emphasize these

hydrophobic regions. Amylose, a principle component of starch, consists of α(14)

glycosidic linkages between glucose monomers, and is generally found in a tight helix

structure, which is largely insoluble in water. Similar to the amylose helix is cyclodextrin,

which has a relatively hydrophobic inner surface that can trap guest molecules. Cellulose,

which consists of β(14) glycosidic linkages, also forms insoluble fibers. In contrast,

dextran is a naturally occurring glucan with high water solubility. The orientation and degrees

of freedom conferred by its relatively flexible α(16) linkages hinder the formation of stable

lamellar structures by the chains.

Dextran is not completely flexible, however, with a persistence length exceeding 2.75 nm.8

Nor is it completely free of ordered association, sometimes aggregating to form flakes in

aqueous solutions.9 Inter-chain associations have been shown in solutions of high dextran

concentrations,10

and evidence of crystallinity in dextran solutions has been shown following

treatment with ethanol.11

As with any polymer, these associations depend on chain length,

polydispersity, and the presence of side groups. Dextran exhibits a certain degree of

branching (about 5%), with predominantly (13) side chains, about 85% of which are one or

two units long and 15% of which are long (30+ units).12

1.3 Polymer brushes

The term “polymer brush” describes a system of polymer chains end-attached to a surface.

Conventional usage of the term implies the presence of one or a few polymer species with a

Page 12: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

11

suitably low polydispersity, in contrast to the diversity of the extracellular membrane, for

example. The chain length must generally exceed the radius of gyration (Rg) of the species in

the medium under study, in contrast to a self-assembled monolayer. Finally, the polymer

chains must be attached with sufficient density for inter-chain interactions to force them into

an extended conformation, stretching away from the surface, in the presence of a good

solvent. Systems that satisfy only the latter criteria can be termed “brush-like”. Conversely,

systems that satisfy all but the last criteria, such as low-density end-attached polymer

coatings or adsorbed complex molecules such as GAGs, may also be described as “brush-

like”.

Polymer and polyelectrolyte brushes in an aqueous environment have been shown to exhibit

ultra-low friction under sliding conditions, which has been attributed to brush resistance to

interpenetration, as well as hydration of the films.13-16

The mechanisms that govern interfacial

sliding of solvated polymer brush-covered surfaces create a unique system distinct from

classical tribological or rheological pictures. The polymer chains, forced into an extended

conformation by interaction with their neighbors and the solvent, resist compression by a

counterface or penetration by asperities or an opposing polymer brush, while maintaining a

fluid layer at the interface even under moderate compression.17

At low loads, the friction

between brush-coated surfaces is below the detection limit of laboratory techniques.17

With

increasing load, the lubricity of opposing brush-coated surface is reduced stepwise as

different mechanisms come into play.18,19

There are several routes to forming a polymer brush on a surface. A “grafting-to” approach

typically exposes the surface to a solution containing fully formed polymer chains with

reactive ends, which diffuse to the surface and attach there. This route is versatile with

respect to chain chemistry, and allows the chains to be characterized prior to adsorption.

When incorporating long chains (relative to their Rg), grafting-to brushes tend to suffer from

low chain densities, since the brush layer itself forms a barrier against chain diffusion to the

surface. The alternative “grafting-from” approach exposes the surface to a solution of

monomers, and polymerization is initiated at the surface. This route can yield the highest

chain densities, but is least versatile with respect to chain chemistry. Specifically,

polysaccharide chains such as dextran must be synthesized enzymatically, and so cannot be

produced by polymerization from a monomer solution.

Page 13: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

12

One approach to increasing chain density without limiting chain selection is the “graft

copolymer” approach, where multiple chains are grafted to a polymer backbone in solution,

and the resulting copolymer is then adsorbed onto the surface. This route is the most versatile

with respect to chain chemistry, can be characterized prior to adsorption, and can also yield

high brush densities compared with most other grafting-to routes. One predominant example

of this is poly(L-lysine)-graft-poly(ethylene glycol) (PLL-g-PEG),20-23

a bottlebrush

copolymer consisting of PEG chains grafted onto the backbone of PLL, a polycation bearing

positively charged amine residues. The resulting graft copolymers are highly hydrophilic, and

the positive charges of the ungrafted amines cause the backbone to spontaneously adsorb

onto negatively charged surfaces, such as metal oxides or mica, while the PEG side chains

extend outwards into solution. The appropriate choice of backbone length, side chain length,

and ratio of grafted to ungrafted lysine residues ensure that the molecules graft with sufficient

density for the side-chains to form a brush. The adsorbed density can be controlled by

varying the immersion time, and density gradients on a surface can be produced using a

computerized dipping system. A similar molecule bearing polysaccharide side chains, PLL-g-

dextran,24-27

is the primary object of this study.

Another route to achieve yet higher density, at further expense of brush quality, is to spin-

coat a reactive surface with polymer chains and then immobilize them to the surface,

producing a polydisperse loop-train-tail structure. One example of this method utilizes the

graft copolymer poly(allylamine)-graft-perfluorophenylazide (PAAm-g-PFPA),28,29

in which

another polycationic backbone, PAAm, is partially functionalized with PFPA groups. The

unreacted positive residues spontaneously adsorb onto negatively charged surfaces. After

adsorption of PAAm-g-PFPA onto a surface, a solution containing the desired polymer

(dextran, in the case of this study) is spin-coated onto a surface and dried. Then, a UV light or

elevated temperature is used to activate the azides, which bind randomly to the polymer (by

the hydroxyl residues in the case of dextran). The resulting chain density can be significantly

higher than those achieved in the case of graft copolymers.29

However, the polymers are

unlikely to be end-attached, and in fact may be attached in multiple places, forming loops and

tails.

Page 14: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

13

1.4 Scope of the thesis

The aim of this project is to determine how the structure and environment of surface-attached

dextran systems influence their hydration and lubricity. The ultimate goals are to gain insight

into the lubricating properties of polysaccharides brushes, thereby helping to illuminate the

behavior of their natural counterparts.

In two preceding projects, the lubricity and hydration of surface-attached polysaccharide

systems in aqueous conditions were investigated. In one project, the copolymer PLL-g-

dextran was synthesized and adsorbed onto substrates to form dextran brushes, and the

lubricity and non-fouling properties of these were compared to those of PLL-g-PEG.26,27

In

the other project, different glycoproteins from synovial fluid were adsorbed onto surfaces and

studied. Those systems were also shown to reduce friction.7 Both projects also studied the

hydration of their respective systems, with the finding that the systems were highly hydrated

but that higher hydration was not necessarily correlated with lower friction. The combined

results indicated that the lubricity and hydration of surface-attached polysaccharides were

worthy of further study, which was the origin of the present project.

This thesis focuses on dextran-based brushes, which are a desirable model system because of

their relatively well-defined composition and conformation. Friction studies of PLL-g-

dextran brushes explore the influence of chain density and applied load, as well as the role of

additives that enhance or impair hydrogen bonding between the chains. The influence of

dextran chain density and brush conformation is further explored by comparison to a PAAm-

g-PFPA-dextran system. The hydration of PLL-g-dextran brushes is investigated by different

experimental methods as well as via computer simulations. At the end of the project, a better

understanding of what governs the lubricity of polysaccharide brushes and films under

boundary conditions could be gained.

1.5 References

1. Alison C. Dunn, John A. Tichy, Juan M. Urueña and W. Gregory Sawyer, “Lubrication

regimes in contact lens wear during a blink”, Tribology International, 2013,

10.1016/j.triboint.2013.01.008

Page 15: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

14

2. Duncan Dowson, “Bio-tribology”, Faraday Discussions, 2012, 10.1039/C2FD20103H

3. Monica Ratoi and Hugh A. Spikes, “Lubricating Properties of Aqueous Surfactant

Solutions”, Tribology Transactions, 1999, 10.1080/10402009908982244

4. Pamela E. Prete, Arzu Gurakar-Osborne and Moti L. Kashyap, “Synovial fluid lipids and

apolipoproteins: a contemporary perspective”, Biorheology, 1995, 10.1016/0006-

355X(95)00001-1

5. Bruno Zappone, Marina Ruths, George W. Greene, Gregory D. Jay and Jacob N.

Israelachvili, “Adsorption, Lubrication, and Wear of Lubricin on Model Surfaces:

Polymer Brush-Like Behavior of a Glycoprotein”, Biophysical Journal, 2007,

10.1529/biophysj.106.088799

6. Seunghwan Lee and Nicholas D. Spencer, “Sweet, Hairy, Soft, and Slippery”, Science,

2008, 10.1126/science.1153273

7. Marcella Roba, Marco Naka, Emanuel Gautier, Nicholas D. Spencer and Rowena

Crockett, “The adsorption and lubrication behavior of synovial fluid proteins and

glycoproteins on the bearing-surface materials of hip replacements”, Biomaterials, 2009,

10.1016/j.biomaterials.2008.12.062

8. Sunil K. Garg and Salvatore S. Stivala, “Assessment of Branching in Polymers from

Small-Angle X-Ray Scattering (SAXS)”, Journal of Polymer Science: Polymer Physics

Edition, 1978, 10.1002/pol.1978.180160808

9. Donald E. Cadwallader Jr., Charles H. Becker, Joan H. Winters and David Marcus, “A

note on particulate matter encountered in some dextran injections”, Journal of the

American Pharmaceutical Association, 1958, 10.1002/jps.3030471219

10. Masuo Aizawa, Shuichi Suzuki, Tatuo Kuoka, Noriyasu Nakajima and Yutaka Iwao,

“Stability of Dextran Solutions”, Bulletin of the Chemical Society of Japan , 1976,

10.1246/bcsj.49.2061

Page 16: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

15

11. Allene Jeanes, N. C. Schieltz and Carl Wilham, “Molecular association in dextran and in

blanched amylaceous carbohydrates”, Journal of Biological Chemistry, 1948, 176, 617-

627

12. Anthony N. de Belder, “Dextran”, Amersham Biosciences, 2003

13. Uri Raviv, Suzanne Giasson, Nir Kampf, Jean-François Gohy, Robert Jérôme and Jacob

Klein, “Lubrication by charged polymers”, Nature, 2003, 10.1038/nature01970

14. Uri Raviv, Suzanne Giasson, Nir Kampf, Jean-François Gohy, Robert Jérôme and Jacob

Klein, “Normal and Frictional Forces between Surfaces Bearing Polyelectrolyte

Brushes”, Langmuir, 2008, 10.1021/la7039724

15. Markus Müller, Seunghwan Lee, Hugh A. Spikes and Nicholas D. Spencer, “The

Influence of Molecular Architecture on the Macroscopic Lubrication Properties of the

Brush-Like Co-polyelectrolyte Poly(L-lysine)-g-poly(ethylene glycol) (PLL-g-PEG)

Adsorbed on Oxide Surfaces”, Tribology Letters, 2003,

10.1023/B:TRIL.0000003063.98583.bb

16. Seunghwan Lee, Markus Müller, Monica Ratoi-Salagean, Janos Vörös, Stéphanie

Pasche, Susan M. De Paul, Hugh A. Spikes, Marcus Textor and Nicholas D. Spencer,

“Boundary Lubrication of Oxide Surfaces by Poly(L-lysine)-g-poly(ethylene glycol)

(PLL-g-PEG) in Aqueous Media”, Tribology Letters, 2003, 10.1023/A:1024861119372

17. Jacob Klein, Eugenia Kumacheva, Diana Mahalu, Dvora Perahia and Lewis J. Fetters,

“Reduction of frictional forces between solid surfaces bearing polymer brushes”, Nature,

1994, 10.1038/370634a0

18. Jacob Klein, Eugenia Kumacheva, Dvora Perahia and Lewis J. Fetters, “Shear forces

between sliding surfaces coated with polymer brushes: The high friction regime”, Acta

Polymerica, 1998, 10.1002/(SICI)1521-4044(199810)49:10/11<617::AID-

APOL617>3.0.CO;2-8

Page 17: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

16

19. Benoît Liberelle and Suzanne Giasson, “Friction and Normal Interaction Forces between

Irreversibly Attached Weakly Charged Polymer Brushes”, Langmuir, 2008,

10.1021/la702367f

20. Gregory L. Kenausis, Janos Vörös, Donald L. Elbert, Ningping Huang, Rolf Hofer,

Laurence Ruiz-Taylor, Marcus Textor, Jeffrey A. Hubbell and Nicholas D. Spencer,

“Poly(L-lysine)-g-Poly(ethylene glycol) Layers on Metal Oxide Surfaces: Attachment

Mechanism and Effects of Polymer Architecture on Resistance to Protein Adsorption”,

Journal of Physical Chemistry B, 2000, 10.1021/jp993359m

21. Ning-Ping Huang, Roger Michel, Janos Voros, Marcus Textor, Rolf Hofer, Antonella

Rossi, Donald L. Elbert, Jeffrey A. Hubbell and Nicholas D. Spencer, “Poly(L-lysine)-g-

poly(ethylene glycol) Layers on Metal Oxide Surfaces: Surface-Analytical

Characterization and Resistance to Serum and Fibrinogen Adsorption”, Langmuir, 2001,

10.1021/la000736+

22. Seunghwan Lee and Nicholas D. Spencer, Chapter 21 in Superlubricity, Elsevier Science

B.V., Amsterdam, 2007

23. Seunghwan Lee, Markus Müller, Monica Ratoi-Salagean, Janos Vörös, Stéphanie Pasche,

Susan M. De Paul, Hugh A. Spikes, Marcus Textor and Nicholas D. Spencer, “Boundary

Lubrication of Oxide Surfaces by Poly(L-lysine)-g-poly(ethylene glycol) (PLL-g-PEG) in

Aqueous Media”, Tribology Letters, 2003, 10.1023/A:1024861119372

24. Anwarul Ferdous, Hiromitsu Watanabe, Toshihiro Akaike and Atsushi Maruyama,

Poly(L-lysine)–graft–dextran copolymer: amazing effects on triplex stabilization under

physiological pH and ionic conditions (in vitro), Nucleic Acids Research, 1998, 26, 3949

25. Shoichiro Asayama, Atsushi Maruyama and Toshihiro Akaike, “Comb-Type

Prepolymers Consisting of a Polyacrylamide Backbone and Poly(L-lysine) Graft Chains

for Multivalent Ligands”, Bioconjugate Chemistry, 1999, 10.1021/bc980093y

26. Chiara Perrino, Seunghwan Lee, Sung Won Choi, Atsushi Maruyama and Nicholas D.

Spencer, “A Biomimetic Alternative to PEG as an Antifouling Coating: Resistance to

Page 18: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

17

Non-Specific Protein Adsorption of Poly(L-lysine)-graft-Dextran”, Langmuir, 2008,

10.1021/la800947z

27. Chiara Perrino, Seunghwan Lee and Nicholas D. Spencer, “Surface-grafted Sugar Chains

as Aqueous Lubricant Additives: Synthesis and Macrotribological Tests of Poly(L-

lysine)-graft-Dextran (PLL-g-dex) Copolymers”, Tribology Letters, 2009,

10.1007/s11249-008-9402-6

28. Michele A. Bartlett and Mingdi Yan, “Fabrication of Polymer Thin Films and Arrays

with Spatial and Topographical Controls”, Advanced Materials, 2001, 10.1002/1521-

4095(200110)13:19<1449::AID-ADMA1449>3.0.CO;2-M

29. Ângela Serrano, Olof Sterner, Sophie Mieszkin, Stefan Zürcher, Samuele Tosatti,

Maureen E. Callow, James A. Callow and Nicholas D. Spencer, “Nonfouling Response

of Hydrophilic Uncharged Polymers”, Advanced Functional Materials, 2013,

10.1002/adfm.201203470

Page 19: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

18

Chapter 2: Experimental Methods

This chapter introduces the experimental methods used in this work. As each subsequent

chapter has its own complete Methods and Materials section, only overviews and remarks of

select methods are discussed here.

2.1 Optical characterization of thin films

2.1.1 Principles and models

Measurements of the mass of an adsorbed layer on a substrate can be carried out by optical

means. Light is guided to the layer, and the resulting interaction (transmission or reflection)

induces a change in the phase, amplitude and/or polarization of the light. These parameters

are measured (directly or indirectly) by the instrument, and fit to an optical model suitable for

the geometry of the substrate and the nature of the light, to yield the thickness and/or

refractive index of the layer. These values can then be related to the sample composition to

determine the surface concentration of polymer. In the case of a multilayer system, each layer

must be measured separately if one wishes to distinguish them.

In the case of a truly homogeneous layer of constant density and known refractive index, the

optical thickness represents the actual thickness of the layer, from which a mass logically

follows. However, the effect on light is much more strongly correlated with the amount of

material present than its spatial organization, so for heterogeneous substances that can swell

or collapse spontaneously, such as polymer brushes and hydrogels, the optical thickness

correlates more with the adsorbate mass than with its true thickness. An empirical linear

correlation between the density and refractive index of a medium was first reported by

Gladstone and Dale in 1863,1 and later that finding was reformulated by de Feijter et al. to

pertain to thin films:

, with M being the mass per surface area of polymer, D the

thickness of the polymer, n the refractive index of the polymer, n0 the refractive index of the

medium, and dn/dc the refractive index increment of the polymer in the medium.2

In a liquid medium, the refractive index increment dn/dc is chosen by simply measuring the

refractive index for different polymer concentrations in solution and fitting the results to a

line. In the case of polymer films measured in air, determining the refractive index increment

Page 20: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

19

is less straightforward, since polymers cannot be easily uniformly distributed through bulk

air. Therefore, various workarounds must be employed to calculate the adsorbed mass

directly from measurements in air.3 For example, the optical mass may be calculated from a

measurement in liquid, and then divided by the thickness of a measurement in air.

In optical measurements of films on the order of 10nm or less, the values for thickness and

refractive index are coupled, and the difficulty of resolving their independent values varies

depending on the experimental setup. Therefore, one of the two parameters (usually the

refractive index) may have to be assumed, and the absolute value that results from the fitting

must be treated with that fact in mind. Therefore the calculated thickness assuming this

refractive index cannot be thought of as a physical thickness, but as representing some

combination of thickness and refractive index of the film. In fact for the optical mass models

described above, when measuring films in air the choice of refractive index is not critical, as

the Gladstone-Dale relation implies

. So if the actual dry refractive index were

actually 1.4 instead of 1.45, this would correspond to an 11% error in the measured thickness

or mass.

One unfortunately common mistake is to perform an optical measurement of a thin film, and

the corresponding data analysis, using a default refractive-index setting, calculate the optical

mass with the de Feijter equation assuming the same refractive index, and then later decide

that a different refractive index may be more suitable for the film, changing the value in the

de Feijter equation to recalculate the mass without redoing the data analysis of the original

optical data with the new refractive index assumption. One must keep in mind that the

thickness has not been measured directly, but rather calculated from the optical parameters

measured using the assumed refractive index – and therefore must be recalculated if one

wishes to use a different refractive index.

2.1.2 Variable-angle spectroscopic ellipsometry (VASE)

Ellipsometry is the measurement of phase and amplitude shifts in polarized light upon

reflection from a set of interfaces. Specifically, the complex ratio of the reflection

coefficients of the different light polarizations is measured as the light interacts with a

multilayered substrate,4 as shown in Figure 2.1. Spectroscopic ellipsometry also measures the

dependence of these parameters on the wavelength of light. VASE is typically performed on

silicon wafers under ambient lab conditions, in air or inside a liquid cell.

Page 21: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

20

Figure 2.1 Representation of polarized light reflecting off of a biomolecule-coated silicon

wafer. The polarized light beams are attenuated and phase shifted by interaction with the

three-layered sample.

2.1.3 Optical waveguide lightmode spectroscopy (OWLS)

Optical waveguide lightmode spectroscopy is an optical biosensing technique that monitors

the grating-assisted in-coupling of a He-Ne laser into the planar waveguide.5 Total internal

reflection guides the light to the end of the waveguiding layer, where it is detected by a

photodiode. The critical angle of reflection is related to the change in refractive index near

the waveguide surface upon adsorption of molecules. From this information, and the

refractive index increment (dn/dc), the mass and refractive index of the adsorbed layer are

calculated by the software. The optical waveguide chip is composed of Si0.75Ti0.25O2 on glass,

presumably with < 10 nm RMS roughness.

2.1.4 Transmission interference adsorption sensing (TInAS)

Transmission interferometric adsorption sensing (TInAS) is an optical biosensing technique

that monitors interference fringes formed when light is transmitted through a film adsorbed

onto a special multilayered TInAS sensor (8 nm RMS roughness with some 10-100 nm

grains), as shown in Figure 2.2. Measurements are made in real time with a sensitivity on the

order of 1 ng/cm2 (when converted to adsorbed density) and a temporal resolution below 1 s.

6

The wavelengths of the interference peaks are related to the optical thicknesses of the layer,

and the default software measures this thickness using an assumed refractive index. However,

the TInAS-FIT software package can use the peak shapes to also fit the refractive index.7

Page 22: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

21

While the TInAS may be measured in air, in practice it is more suited to measurements in

liquid, as the polymer may be adsorbed onto the surface in situ without disturbing the setup

from the baseline measurement, and subsequent drying induces instrumental drift, which is

problematic for such thin films.

Figure 2.2 Illustration of the operating principle of TInAS. White light passes through a

custom multilayer sensor, and multiple partial reflections interfere to produce modulation of

the transmitted light intensity with wavelength.

2.2 Measurement of friction with the atomic force

microscope

Atomic force microscopy (AFM) is widely used for topographic imaging and force

spectroscopy, and may also be used for nanotribological studies via a technique called lateral

force microscopy. A probe, in this case consisting of a borosilicate colloidal sphere glued to

the end of a cantilever, is scanned laterally along a surface, and the normal and lateral

deflections are measured independently. Given the known properties of the cantilever, these

can in principle be calibrated to measure normal and lateral (friction) force in real time,

although not all AFMs allow for the quantitative calibration of lateral forces. The lateral force

is obtained by scanning the probe perpendicularly to the cantilever and recording the TMR

(Trace Minus Retrace in Volts) value. The TMR value is proportional to the friction force and

minimizes the contributions to the lateral force from non-friction sources,8 but both the

Page 23: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

22

friction force Ff and local coefficient of friction µ=dFf/dFL, FL being the applied load, are

given in arbitrary units. Useful friction comparison may nevertheless be obtained by making

measurements on a heterogeneous sample in a single run with the same tip and cantilever, the

resulting values being directly comparable with each other. Calibration of the deflections

(normal or lateral) is achieved by measuring the spring constant of the cantilever (prior to

colloid attachment!) as well as the ratio of photodiode response in Volts to deflection of the

cantilever in meters (at the time of measurement!). This is commonly known as the InvOLS

value for normal force measurements.

An important consideration during measurement is the positioning of the laser spot on the

cantilever. The laser spot typically has a large size compared to the width of the cantilever,

and some of the light is lost when focusing near the tip of the cantilever. The measurement

sensitivity (inverse of the InvOLS) depends on the strength of the laser signal measured by

the photodiode (the sum) as well as the distance from the end of the deflecting cantilever, as

shown in Figure 2.3.

Page 24: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

23

Figure 2.3 Factors influencing the measurement sensitivity. Focusing the laser spot too near

the cantilever tip will reduce the intensity, while positioning it too far away will reduce the

influence of cantilever bending.

Varying the scan size is not likely to make an experimental difference in non-wearing

experiments unless the scan size is reduced to sizes on the order of the contact area with the

probe, which can change the nature of the test. In contrast, varying the scan rate is likely to

change the measured friction by sampling dissipative forces on a different timescale. Varying

the normal force may influence the contact area, and therefore change the influence of the

probe morphology (roughness) on the experiment. Varying the probe size can have similar

effects, as colloidal spheres of different sizes tend to have different surface roughnesses. An

important consideration when selecting the cantilever and colloidal probe is to minimize in-

plane bending,9 which is an alternative mechanical response to lateral forces that, unlike

twisting, cannot be measured by the laser/photodiode setup. Larger probes and shorter

cantilevers are optimal for minimizing the ratio of in-plane bending to twisting in response to

friction forces.

2.3 Remarks on computer simulations

While computer simulations become increasingly widespread in the study of biomolecules,

there often exists an unfortunate divide between experimentalists and modelers. In the

interest of making the topic more accessible to experimentalists, it should be pointed out that,

in most cases, the accurate determination of the equilibrium state of a complex molecule is a

far simpler task than modeling physically accurate kinetic processes. The kinetics of a ball

rolling down a hill can be difficult to predict correctly if one attempts to model terrain

roughness, friction, sticky soil, and air resistance. In contrast, determining that it must

eventually end up at the bottom of the hill can be as simple as a couple of potential energy

calculations and a ballpark estimation of dissipative forces. Therefore the distinction between

studies of equilibrium conformations and studies of kinetic processes should be kept in mind

when trying to understand any simulation work.

The other difficulty inherent to course-grained simulations is the selection of useful

parameterization of the forces in question. The work described in Chapter 6 used a course-

Page 25: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

24

grained model of dextran in water from the literature, which parameterized all non-

neighboring interactions using the standard Morse equation – therefore neglecting many finer

aspects of hydrogen bonding, such as directional heterogeneity or cooperative effects.10

While this was deemed sufficiently accurate to determine the overall brush structure, such

course graining inevitably imposes limits on the fidelity of the resulting system, which should

be kept in mind during any discussion or comparison to other findings.

2.4 References

1. John H. Gladstone and T. P. Dale, “Researches on the Refraction, Dispersion, and

Sensitiveness of Liquids”, Philosophical Transactions of the Royal Society of London,

1863, 10.1098/rstl.1863.0014

2. J. A. De Feijter, J. Benjamins and F. A. Veer, “Ellipsometry as a tool to study the

adsorption behavior of synthetic and biopolymers at the air–water interface”,

Biopolymers, 1978, 10.1002/bip.1978.360170711

3. Tao Wu, Peng Gong, Igal Szleifer, Petr Vlcek, Vladimír Šubr and Jan Genzer, “Behavior

of Surface-Anchored Poly(acrylic acid) Brushes with Grafting Density Gradients on

Solid Substrates: 1. Experiment”, Macromolecules, 2007, 10.1021/ma0710176

4. Harland G. Tompkins, “Spectroscopic Ellipsometry and Reflectometry”. Wiley, New

York. 1999

5. Janos Vörös, J. J. Ramsden, G. Csucs, I. Szendro, Markus Textor and Nicholas D.

Spencer, “Optical grating coupler biosensors”, Biomaterials, 2002, 10.1016/S0142-

9612(02)00103-5

6. Manfred Heuberger and Tobias E. Balmer, “The transmission interferometric adsorption

sensor”, Journal of Physics D: Applied Physics, 2007, 10.1088/0022-3727/40/23/S07

7. Takumi Sannomiya, Tobias E. Balmer, Manfred Heuberger and Janos Vörös,

“Simultaneous refractive index and thickness measurement with the transmission

interferometric adsorption sensor”, Journal of Physics D: Applied Physics, 2010,

Page 26: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

25

10.1088/0022-3727/43/40/405302

8. Ernst Meyer, “Atomic Force Microscopy”, Progress in Surface Science, 1992,

10.1016/0079-6816(92)90009-7

9. John E. Sader and Christopher P. Green, “In-plane deformation of cantilever plates with

applications to lateral force microscopy”, Review of Scientific Instruments, 2004,

10.1063/1.1667252

10. Valeria Molinero and William A. Goddard III, “M3B: A Coarse Grain Force Field for

Molecular Simulations of Malto-Oligosaccharides and Their Water Mixtures”, Journal of

Physical Chemistry B, 2004, 10.1021/jp0354752

Page 27: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

26

Chapter 3: Load-induced transitions in the

lubricity of adsorbed poly(L-lysine)-g-

dextran as a function of polysaccharide

chain density

This work was published in 2011 in ACS Applied Materials and Interfaces.50

The

supplementary materials have been added to the end of this chapter. The major part of this

work was done by myself and Dr. Kenneth Rosenberg working together, with some initial

friction measurements made solely by him and some final friction measurements and

calculations made solely by myself. Dr. Rowena Crockett and Prof. Nicholas D. Spencer

supervised the work, contributed in discussions about the project steps, and corrected the

manuscript.

3.1 Abstract

Chain-density gradients of poly(L-lysine)-graft-dextran (PLL-g-dex), a synthetic comblike

copolymer with a poly(L-lysine) backbone grafted with dextran side chains, were fabricated

on an oxidized silicon substrate. The influence of the changing dextran chain density along

the gradient on the local coefficient of friction was investigated via colloidal-probe lateral

force microscopy, as shown in Figure 3.1. Both in composition and structure, PLL-g-dex

shares many similarities with bottlebrush biomolecules present in natural lubricating systems,

while having the advantage of being well-characterized in terms of both architecture and

adsorption behavior on negatively charged oxide surfaces. The results indicate that the

transition of the dextran chain density from the mushroom into the brush regime coincides

with a sharp reduction in friction at low loads. Above a critical load, the friction increases by

more than an order of magnitude, likely signaling a pressure-induced change in the brush

conformation at the contact area and a corresponding change in the mechanism of sliding.

The onset of this higher-friction regime is moved to higher loads as the chain density of the

film is increased. While in the low-load (and low-friction) regime, increased chain density

leads to lower friction, in the high-load (high-friction) regime, increased chain density was

found to lead to higher friction.

Page 28: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

27

Figure 3.1 Gradients of PLL-g-dextran were fabricated on silicon wafers by controlled

dipping, and the influence of the changing dextran chain density on the local coefficient of

friction was investigated via colloidal-probe lateral force microscopy.

3.2 Introduction

The molecular architecture and structure of bottle-brush polymers such as poly(l-lysine)-

graft-poly(ethylene glycol) (PLL-g-PEG), and their use in modifying surface properties, have

been the subject of several recent studies.1-5

In particular, these polymers are attractive

friction-reduction additives in an aqueous environment, because they tend to adsorb onto

surfaces to form dense brushlike layers, in which the highly solvated nature of the extending

polymer chain is known to correlate well with effective lubrication.3,4,6-8

Recently, the

lubrication properties of poly(L-lysine)-graft-dextran (PLL-g-dex) with a variety of grafting

ratios and dextran molecular weights under aqueous conditions were also studied both on the

macro- and the nanoscale.9,10

It was shown that PLL-g-dex lubricates metal oxide surfaces

very efficiently, and readily forms high-density polymer brush layers with properties

dependent on the molecular weight of the dextran side chains.

Dextran is a naturally occurring polysaccharide consisting of an α(1→6)-linked glucan with

branches attached to the 3-positions of the straight-chain glucose units. PLL-g-dex is formed

by grafting dextran chains onto the backbone of poly(L-lysine), a polycation that

spontaneously adsorbs onto negatively charged surfaces, such as metal oxides, at multiple

points via the ungrafted, primary amine groups, as shown in Figure 3.2. The resulting graft

copolymer, PLL-g-dex, is highly soluble in water, and, when bound to a substrate at

Page 29: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

28

sufficient density, effectively renders the surface nonfouling, similarly to other dextran-

containing surfaces.9,11-15

The overall stability of PLL-g-dex, similar to that of PLL-g-PEG, is

a feature that biomedical applications can readily exploit, for example as a gene-delivery

vector.16,17

In contrast to PEG, however, dextran is not prone to undesirable oxidation

reactions and possesses multiple reactive sites, enabling further functionalization and

chemical modification.

Figure 3.2 Structure of PLL-g-dextran copolymer with grafting ratio k + 1.

An important characteristic of PLL-g-dex that it shares with virtually all lubricious

biomolecules is the presence of oligosaccharide side chains. Many studies have focused on

the lubricating properties of glycoproteins, including lubricin and naturally occurring

polysaccharides on a variety of substrates and under various environmental conditions.18

Low

friction and wear are normally associated with the complex mixtures that constitute natural

lubricious layers, which often consist of highly hydrated glycoproteins, such as proteoglycans

and mucins.18

A common difficulty in interpreting these measurements lies in the complexity

of the structure and composition of the glycoproteins. The topological complexity of the

molecules, and the subtle differences in the electrostatic and hydrophobic interactions that

govern the adsorption properties, lead to an uncertainty in adsorbed conformation. This

presents a challenge when attempting to determine the features contributing to lubricity. In

Page 30: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

29

PLL-g-dex, the composition and structure are much simpler than in glycoproteins, and thus it

constitutes a biomimetic system that maintains the critical features important for lubricity.

In the present study, we examine the relationship between the density of dextran chains and

the frictional properties of the resulting system. Although a great deal is known regarding the

structure of polymer films and their interactions, much less is understood regarding how

changes in polymer conformation and, as a consequence, film mechanical properties,

influence friction.19-21

The conformation of surface-attached polymer chains is dependent on

the mean spacing between neighboring chains L and the radius of gyration, Rg. When L >>

2Rg, the non-overlapping, isolated chains are restricted in conformation only by the surface,

and may assume a pancake- or mushroom-shaped conformation, depending on the interaction

of the polymer with the surface. At the other extreme, for L < 2Rg (the brush regime), the

interchain steric repulsive interactions cause the chains to assume more extended

conformations. The use of a gradient to generate gradual changes in polymer chain density is

convenient for producing the full range of conformations. Gradients, in general, are a

powerful technique in studying parametric dependency, as evidenced by their use in a variety

of applications, including wetting22,23

and cell-adhesion24

studies. For example, using a

grafting-density gradient of poly(acrylamide) (PAAm) on silica, Genzer and co-workers

measured the mushroom-to-brush transition point (0.065 chains/nm2), based on height

measurements obtained by ellipsometry, and correlated variations in chain density with

wetting properties.25

3.3 Materials and Methods

3.3.1 Poly(L-lysine)-graft-dextran (PLL-g-dex) The PLL-g-dex copolymer used in these experiments was synthesized as described

previously.9,10

Briefly, dextran (dextran T5, 5.2 kDa, polydispersity 1.8, Pharmacosmos A/S,

Denmark) undergoes reductive amination with poly(L-lysine)-HBr (20 kDa, polydispersity

1.1, Sigma-Aldrich, Switzerland). The solvent for the reaction was a sodium borate buffer

(0.1 M, pH 8.5, 0.4 M NaCl). The terminal dextran aldehyde group was first reacted with the

primary amine groups of PLL to form a Schiff base, which was subsequently reduced by

sodium cyanoborohydride (NaBH3CN). The resultant PLL-g-dex graft copolymer was

separated from the unreacted materials through filtration with ultracentrifugation. The

specific ratio of lysine monomers to dextran chains is defined as the grafting ratio and can be

Page 31: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

30

controlled during synthesis by adjusting the reactant ratio.26,27

The grafting ratio was

determined through 1H-NMR spectroscopy of PLL-g-dex in D2O and elemental analysis

(EA).

The particular copolymer studied here was PLL(20)-g[5.3]-dex(5), where 20 kDa is the molar

mass of PLL including the Br- counterions, 5.3 is the grafting ratio, and 5 kDa is the weight-

average molecular weight (Mw) of the dextran. On average, dex(5) consists of 32 monomer

units, which assuming each monomer is 0.7 nm long corresponds to a fully stretched chain

length of 22 nm.

3.3.2 Gradient preparation Bare silicon wafers of 2 cm × 1 cm were cleaned by sonicating twice in toluene, then twice in

2-propanol, each for five minutes. Next, the substrates were exposed to O2 plasma for two

minutes (Plasma Cleaner/Sterilizer, PDC-32G instrument, Harrick, Ossining, NY, USA). The

thickness of the surface SiO2 layer was measured by ellipsometry (described below), after

which each substrate was immediately dipped at a controlled rate into a 0.02 mg/mL solution

of PLL-g-dex in 1 mM 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid (HEPES) at pH

7.4.28

After removal from the PLL-g-dex HEPES solution, the substrate was first rinsed with

polymer-free HEPES buffer to remove any weakly bound or nonadsorbed PLL-g-dex, and

then copiously rinsed with ultrapure water (Milli-Q Gradient A10, Millipore SA, Molsheim,

France) and finally dried under N2. The final PLL-g-dex thicknesses were then measured in 2

mm increments along the substrate by ellipsometry.

Both SiO2 and PLL-g-dex film thicknesses were measured in ambient air by ellipsometry

using a M-2000 V spectroscopic ellipsometer (J.A. Woollam Co., Inc., Lincoln, NE, USA) at

65°, 70°, and 75° at wavelengths between 370-1000 nm. The variable-angle spectroscopic

ellipsometry (VASE) spectra were analyzed using WVASE32 software. The “dry” thickness

of PLL-g-dex was obtained by fitting the VASE spectra to a multilayer model based on the

optical properties of a generalized Cauchy layer (A = 1.45, B = 0.01, C = 0).29

The results of

ellipsometric characterization of a PLL-g-dex surface-chemical gradient are shown in Figure

3.3A. The dry thickness of the copolymer layer ranges from below the experimental scatter of

SiO2 thickness (0.1 Å) at the bare end to 16 ± 1 Å, corresponding to the maximum thickness

for this particular PLL-g-dex architecture.

Page 32: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

31

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

0 5 10 15 20

Dry

he

igh

t, Å

Distance from end of substrate, mm

(A)

0

0.5

1

1.5

2

2.5

3

3.5

4

0 5 10 15 20

L/2

Rg

Distance from end of substrate, mm

(B)

Page 33: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

32

Figure 3.3 (A) “Dry” height of PLL-g-dex along a SiO2 substrate measured using

ellipsometry. (B) Plot of L/2Rg along the gradient obtained by calculating the mean spacing

between dextran chains. The dashed line is not a fit of the data, but rather is calculated

directly from relating the PLL-g-dex adsorption time to adsorbed mass using OWLS data. (C)

The resulting chemical gradient consists of isolated dextran chains at one end (L >> 2Rg) and

a fully stretched “brush” regime at the opposite end (L < 2Rg).

The adsorbed mass of PLL-g-dex onto a standard Si0.75Ti0.25O2 waveguide in buffer solution

was measured via optical waveguide lightmode spectroscopy (OWLS), an optical biosensing

technique which monitors the grating-assisted in-coupling of a He–Ne laser into the planar

waveguide. Total internal reflection guides the light to the end of the waveguiding layer,

where it is detected by a photodiode. The critical angle of reflection is related to the change in

refractive index near the waveguide surface upon adsorption of molecules. From this

information, and the refractive index increment (dn/dc), the mass and refractive index of the

adsorbed layer are calculated30-32

by the software. The refractive index increment of PLL-g-

dex was taken to be 0.131.9

Prior to the adsorption experiment, the optical waveguide chip (standard: Si0.75Ti0.25O2 on

glass, Microvacuum, Budapest, Hungary) was cleaned by ultrasonication in 0.1 M HCl for 10

min, rinsed with ultrapure water, ultrasonicated in 2-propanol for 10 min, rinsed again with

ultrapure water, dried under nitrogen, and cleaned in a UV/ozone cleaner (UV/Clean, model

135500, Boeckel industries Inc., Feasterville, PA) for 30 min. The cleaned waveguide was

placed into the OWLS flow cell and equilibrated by exposing to HEPES buffer solution

overnight, in order to obtain a stable baseline. The waveguide was then exposed to a polymer

solution (0.015 mg/mL in a similar HEPES buffer) under constant flow for at least 30 min,

resulting in the formation of a polymer adlayer.

3.3.3 Atomic force microscope measurements

The AFM experiments were conducted in contact mode using a Dimension 3000 AFM

(Nanoscope III controller, Digital Instruments, Santa Barbara, CA, USA) and a MFP-3D

AFM (Asylum Research, Santa Barbara, CA, USA). The probes were commercially available

V-shaped Si3N4 cantilevers (NP-S20, nominal normal spring constant 0.58 N/m, Veeco,

Page 34: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

33

Plainview, NY, USA) and rectangular Si cantilevers (CSC12A, nominal normal spring

constant 0.95 N/m, Mikromasch, Tallinn, Estonia).

A borosilicate colloidal sphere (Kromasil®, Eka Chemicals AB, Bohus, Sweden; radius ≈ 3

μm) was glued to the end of the cantilever with UV-curable epoxy (Norland Optical

Adhesive #61 or #63, Norland, Cranbury, NJ, USA). The entire assembly was then irradiated

with UV light for 30 min and allowed to sit for at least 1 day to further strengthen the

adhesion between the colloidal sphere and the cantilever. The nanotribological experiments

described here include measurements taken first with a bare, uncoated SiO2 colloid probe and

then with a PLL-g-dex brush-like layer attached to the same colloidal sphere by the same

method as used for the gradient preparation, and with sufficiently long exposure to solution to

ensure maximum possible brush density (according to the gradient and OWLS kinetics

measurements).

Friction measurements were obtained by scanning the probe perpendicularly to the major axis

of the cantilever and recording the TMR (Trace Minus Retrace in mV) value. The TMR value

is directly proportional to the friction force and minimizes the contributions to the lateral

force from non-friction sources.33

Each bare colloid was first slid against the uncoated part of

the gradient until the coefficient of friction reached a constant value. This “running in” step

was done to check the quality of the tip and to ensure that any transient colloidal wear or

damage could not affect the measured friction. To test for any such changes as well as

material transfer from the PLL-g-dex gradient, the bare surface was scanned intermittently to

ensure the friction response remained constant. Additionally, the gradient was measured in a

non-sequential order to minimize possible effects arising from drifting of force sensitivity.

All scanning distances were 1 μm with velocities of 1 μm/sec, unless otherwise noted. The

dependence of scanning distance on lateral force was investigated and showed that for

lengths over 100 nm, the lateral force remained constant.

The normal and lateral forces were calibrated as described in the Supporting Information.

Briefly, the cantilever spring constants were calibrated from the power spectral density of the

thermal noise fluctuations34

and by the method of Sader and coworkers.35

The lateral force

measurements of the rectangular cantilevers were calibrated using the test probe method

described by Carpick and coworkers,36

and the measurements of the V-shaped cantilevers

were scaled to the resulting curve. Rectangular cantilevers were chosen that approached the

Page 35: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

34

normal spring constant of the V-shaped cantilevers while minimizing errors due to in-plane

bending.37

Following the nanotribological measurements with an uncoated SiO2 colloid, the entire tip-

cantilever assembly was gently lifted off the gradient surface and immediately exposed to a

0.02 mg/mL PLL-g-dex solution in 1 mM HEPES to form a brush layer on the colloid, as

predicted both by the OWLS and ellipsometry measurements. After 30 minutes, the AFM

cantilever was thoroughly rinsed with polymer-free HEPES buffer and placed back into

contact with the same gradient substrate to resume the friction experiments.

3.4 Results

3.4.1 Measurement of L/2Rg

In order to understand how lubrication varies with dextran chain density and conformation,

the parameter L/2Rg is calculated along the gradient. Depending on the specific technique

used, literature values of Rg of 5 kDa dextran range from 1.9 nm (dynamic light scattering)38

to 2.4 nm (fluorescence correlation spectroscopy).39

Because the focus of this study is on the

relative, rather than absolute, change in L/2Rg, the value of Rg throughout this paper is simply

fixed at 1.9 nm. The only effect of setting Rg to 2.4 nm would be to shift all L/2Rg values by

about 25% in the direction of lower chain density. Two independent methods to measure L

are discussed here. First, from the “dry” PLL-g-dex thickness h (Figure 3.3A), the grafting

density of dextran chains σ ≈ 1/L2 can be found from σ = hρNA/MN, where ρ is the bulk

polymer mass density, NA is Avogadro’s number, and MN is the polymer molecular weight.40

Previous adsorption results of PLL-g-dex on SiO2 substrates measured with optical

waveguide lightmode spectroscopy (OWLS) give the maximum amount of adsorbed PLL-g-

dex as 290 ± 16 ng/cm2.9,10

Finally, because h is proportional to σ = 1/L2 from the above

relationship, L can be calculated at each point along the gradient where h is known. These

data points are shown in Figure 3.3B. Although one end of the substrate was not exposed to

PLL-g-dex solution, L remains finite because of the experimental uncertainty in the

ellipsometric measurement, as mentioned in Materials and Methods.

A separate technique to calculate L utilizes the kinetic adsorption data obtained from OWLS

data to map the time exposed to PLL-g-dex solution directly to adsorbed polymer mass. In

this method, there is no assumption in relating dextran chain density to “dry” thickness.

Page 36: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

35

Instead, the adsorption time is known for each position along the gradient, which is then

mapped to the OWLS adsorption curve to generate PLL-g-dex mass as a function of position.

The data from this method is shown as a dashed line in Figure 3.3B, in excellent agreement

with the analytical model described above.

3.4.2 Friction measurements

Figure 3.4 shows three representative “friction-load” curves measured with a bare, uncoated

SiO2 colloid probe at different values of L/2Rg along a PLL-g-dex gradient. The friction-load

curves measured with a coated SiO2 sphere show a similar shape to those in Figure 3.4 (not

shown). The lateral force (FL) increases as a function of normal force (FN) and initially

displays linear behavior, as seen in the figure inset. However, upon increasing the load past a

critical value, the linear relationship ceases to be valid as the friction, μ=dFL/dFN, rapidly

increases for each of the curves shown. Continued ramping up of load again results in a

constant, but higher, friction value. This transition was not found to occur when the bare

silica colloid was sliding against a bare silicon wafer, which yielded a stable friction

coefficient of 0.22 across the entire load range, in excellent agreement with literature.41

The

coefficients of friction of both regimes (termed “low load" and “high load") were recorded

along the gradient, i.e., at different values of L/2Rg, and are shown in Figure 3.5. The initial

slope (low load) for both uncoated and coated colloid is displayed in Figure 3.5A while both

slopes (low and high load) are shown in Figure 3.5B. Although the coated colloid exhibits a

lower friction coefficient at low load, it gives a much higher friction coefficient than the

uncoated colloid at high load. The transition in normal load between the low and high friction

regimes, defined as FT, is shown in Figure 3.6.

Page 37: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

36

Figure 3.4 Set of typical friction vs. load curves (A) of a bare SiO2 colloidal probe sliding on a PLL-

g-dex-coated SiO2 gradient for L/2Rg = 2.9 (◊), L/2Rg = 1.2 ( ), and L/2Rg = 0.6 (▲), with (B)

showing the same data zoomed in on the region up to FL = 50 nN.

0

100

200

300

400

500

0 50 100 150 200 250 300

Late

ral f

orc

e, n

N

Normal force, nN

(A)

0

10

20

30

40

50

60

0 10 20 30 40 50

Late

ral f

orc

e, n

N

Normal force, nN

(B)

Page 38: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

37

Figure 3.5 Coefficient of friction plotted against L/2Rg (A) in both low and high load (FL >

100 nN) and (B) just low load (FL < 50 nN) regimes. The squares represent a bare SiO2

colloid at low load (empty) and high load (filled). The circles represent a PLL-g-dex-coated

SiO2 colloid at low load (grey) and high load (black).

0

1

2

3

4

5

6

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

Co

eff

icie

nt

of

fric

tio

n

L/2Rg

(A)

0

0.1

0.2

0.3

0.4

0.5

0.6

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

Co

eff

icie

nt

of

fric

tio

n

L/2Rg

(B)

Page 39: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

38

Figure 3.6 Characteristic load at which the friction starts to deviate from linearity, denoted as

the transition load, is shown as a function of L/2Rg along the gradient with uncoated colloid

sphere (open) and PLL-g-dex-coated colloid sphere (filled).

3.5 Discussion

3.5.1 Effect of L/2Rg

Polymer-brush layers at a tribological interface are known to cause a significant reduction in

the friction coefficient.2-4,42

Essentially, the long-range steric interaction arising from the

osmotic repulsion at high chain density provides a suitable boundary film to reduce, if not

eliminate entirely, direct contact of opposing surfaces.43

Therefore, it is not surprising that,

over the moderate pressures generally studied (in the megapascal range or lower), the lower

the value of L/2Rg, or, equivalently, the higher the dextran chain density, the lower the

coefficient of friction. For L > 2Rg, the coefficient of friction is large and relatively

independent of dextran density. In this non-brush regime, the rather inhomogeneous

distribution of dextran chains along the surface is an ineffective barrier against the external

load. In contrast, when L < 2Rg, the higher density produces a more homogeneous brush layer

0

50

100

150

200

250

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

Tran

siti

on

Lo

ad, n

N

L/2Rg

Page 40: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

39

that can give rise to a greater steric repulsion. With increasing dextran density, the thickness

of the film layer is expected to grow, allowing for greater physical separation for a given load

between sliding surfaces. Moreover, as a result of the increasing density of dextran chains,

the total amount of associated water within the layer also increases.9,10

These combined

factors of increased physical separation and increased film hydration are instrumental in

rapidly lowering the coefficient of friction as the chain density transitions along the gradient

from the bare end until the maximum brush density. However, even sub-optimal coverage

still functions as an effective barrier layer between the colloid and bare substrate, as

evidenced by the roughly twofold friction reduction near L 2Rg. This reduction in friction,

even at partial coverage, is consistent with other published results indicating that modifying

the interfacial properties of solid substrates with adsorbed polymers, not necessarily in the

brush regime, still achieves improved lubricity.44-47

For identical values of L/2Rg, the friction coefficient with the brush-coated colloid is

decreased with respect to the uncoated colloid at low loads (Figure 3.5A). Furthermore, it is

apparent that the friction coefficient is highly dependent on the interfacial properties of the

sliding and stationary surfaces individually. Specifically, with the uncoated SiO2 colloid

sliding against the uncompressed brush, the friction is much lower than when the brush-

coated SiO2 colloid is sliding against a nearly bare (i.e., very high L/2Rg) surface, when both

are at low load. Here, even though both systems consist of similar surfaces within the

tribocontact, the surface histories are very different. Presumably the continual shearing of the

brush-coated colloid against the relatively bare substrate has a wearing effect on the brush, in

comparison with the uncoated colloid sliding across the uncompressed, previously

unperturbed brush.

3.5.2 Effect of load

The tribological response of the PLL-g-dex gradient to external load exhibits two distinct

linear regimes over the loading range studied (Figure 3.4) At low loads, the lateral force

increases linearly up to a transition load FT (Figure 3.4 inset), after which it increases with a

higher slope. From Figure 3.6, the value of FT measured along the gradient with an uncoated

SiO2 colloidal sphere remains virtually constant as L/2Rg is decreased, with a value around 60

nN, until L/2Rg reaches unity, FT nearly doubling by L/2Rg = 0.6. The similarity between the

dependence of FT and friction coefficient on L/2Rg is most likely not coincidental–that is, the

stiffer and more hydrated PLL-g-dex caused by the more densely packed dextran chains leads

Page 41: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

40

to both a decrease in friction and an increase in FT. As expected, with the brush attached to

the colloid, the ability of the polymer film within the tribocontact to support an external load

is increased compared with a bare colloid. In other words, the presence of the brushlike layer

on the sliding colloid helps to maintain a lower friction coefficient before the friction abruptly

rises, above FT. For both the coated and uncoated colloidal sphere, as the dextran chain

density on the gradient increases towards full brush, FT rises in response to the growing steric

repulsion of the PLL-g-dex layer.

Interestingly, although this transition in the friction–load curve has been observed in a variety

of polymer brush systems, both neutral and charged,48

the literature provides no answers as to

a possible mechanism. A similar transition in friction with increasing brush compression,

caused by increasing load, has also been predicted for entangled polymer brush systems,21

in

contrast to nonentangled systems. Although the dextran chains studied are likely too short

and stiff to experience entanglement, their interchain hydrogen bonding can have a

comparable effect. The transition load represents the tipping point of an energy balance

between brush compression and brush parting or interpenetration, for the bare and brush-

coated colloidal spheres, respectively. If interchain hydrogen bonds, which scale with brush

density, are assumed to behave comparably to rapidly forming entanglements, they serve as a

hindrance to this transition by resisting the parting of the brush. Additionally, the friction of

the brush-on-brush system at high loads may be particularly exacerbated by the formation

and disruption of new hydrogen bonds between the chains of the interpenetrating brushes. A

simplified Hertzian analysis shows that the contact pressures in the ‘low’ regime are still far

higher than those commonly achieved in physiological conditions.49

For FL ≈ 10 nN, with a

colloid radius of 3.5 μm, the maximum pressure reaches 50 MPa. For comparison, the

typical contact pressures seen in joints are normally only 10% of this value, and many

lubricated physiological tribosystems have far lower contact pressures (e.g., in the eye).

3.6 Conclusions

The lubricating properties of a density gradient of PLL-g-dex copolymer were investigated by

means of colloidal-probe lateral force microscopy. The dominant parameter determining the

coefficient of friction was the relative spacing (L/2Rg) of the surface-grafted dextran chains.

Furthermore, the nanotribological analysis showed that the friction coefficient along the

gradient is highly load dependent. At high contact loads, the lubrication behavior is very

Page 42: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

41

different compared to that at low loads, both for brush-brush and brush-surface contact.

Although higher polysaccharide chain densities lower friction more effectively at low loads,

they actually increase friction in the high-load regime compared to lower chain densities.

Although this study highlights potential design limits in the use of polysaccharide brushes, it

also opens new potential toward high-friction applications of polymer coatings. Although this

study only examined the effect of dextran chain density, other architectural parameters may

be critical in determining the overall friction, including grafting ratio and dextran molecular

weight. These initial studies should prove useful in better designing and tuning PLL-g-dex

and related copolymers for specific water-based lubrication systems.

3.7 Supporting data: calibration of lateral-force

measurements

The normal force spring constant (nN/nm) of each V-shaped cantilever was calibrated from

the power spectral density of the thermal noise fluctuations in air using the Asylum MFP-3D.

The V-shaped cantilevers were then mounted on the Veeco Dimension for lateral force

measurements, which were obtained by scanning the probe perpendicularly to the cantilever

and recording the TMR (Trace Minus Retrace in mV) value, which is directly proportional to

the friction force. Since all experiments are measured within a gradient in a single run with

the same tip and cantilever, the values from the V-shaped cantilevers are directly comparable

with each other, even in the absence of calibration. In order to facilitate comparison with

other studies, calibration was performed (and additional data obtained) by repeating selected

measurements using rectangular cantilevers mounted on the MFP-3D. For these cantilevers,

the normal and lateral force spring constants were calibrated by measuring the thermal-noise

fluctuations of the tipless cantilever in air, then obtaining the cantilever dimensions from

SEM images and using the Sader method to obtain the spring constants. The sphere was

subsequently attached and a full range of friction measurements performed according to the

same procedure as with the V-shaped cantilevers. The lateral sensitivity of the rectangular

cantilevers was inferred by comparison with another cantilever using the test probe method

described by Carpick and coworkers, yielding absolute values for both the friction force Ff

and local coefficient of friction µ=dFf/dFL, FL being the applied load. The test probe

calibration method itself includes correcting factors for differences in cantilever dimensions,

colloidal sphere size, normal and lateral spring constant of the cantilever, and spot intensity

Page 43: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

42

between the experimental and reference setups. Additionally, all the rectangular cantilevers

used for the calibration and measurement were produced in a single batch from the same

wafer, so that any errors stemming from non-ideal behavior of the cantilevers or unaccounted

factors, such as variations in the reflectivity of the cantilever top coatings, would be

consistent across all measurements.

Finally, all the uncalibrated lateral force measurements from the V-shaped cantilever were

multiplied by a single scaling factor (mV/nN) to match the curve of the calibrated

measurements. The scaled results are shown in Figure 3.7. While the calibration proved

highly effective within the range relevant to the low-friction regime, extrapolation to the

high-friction regime inherently leads to a higher error, due to the non-linearity involved in the

twisting of the V-shaped cantilever. The friction results in the high-friction regime should

therefore be regarded as being a) semiquantitative and b) potentially an underestimate of the

true values.

Page 44: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

43

Figure 3.7 The coefficient of friction of (A) a bare SiO2 colloid and (B) a brush-coated SiO2

colloid against a gradient substrate, plotted against L/2Rg in the low-friction regime. The

filled points represent fully calibrated data measured using rectangular cantilevers, while the

empty points represent scaled data measured using V-shaped cantilevers. Error bars are

shown in Figure 3.5.

0

0.1

0.2

0.3

0.4

0.5

0 0.5 1 1.5 2 2.5 3 3.5 4

Fric

tio

n C

oe

ffic

ien

t

L/2Rg

(A)

0

0.1

0.2

0.3

0.4

0.5

0 0.5 1 1.5 2 2.5 3 3.5 4

Fric

tio

n C

oe

ffic

ien

t

L/2Rg

(B)

Page 45: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

44

3.8 References

1. Seunghwan Lee and Nicholas D. Spencer, Chapter 21 in Superlubricity, Elsevier Science

B.V., Amsterdam, 2007

2. Stéphanie Pasche, Susan M. De Paul, Janos Vörös, Nicholas D. Spencer and Marcus

Textor, “Poly(L-lysine)-graft-poly(ethylene glycol) Assembled Monolayers on Niobium

Oxide Surfaces: A Quantitative Study of the Influence of Polymer Interfacial

Architecture on Resistance to Protein Adsorption by ToF-SIMS and in Situ OWLS”,

Langmuir, 2003, 10.1021/la034111y

3. Seunghwan Lee, Markus Müller, Monica Ratoi-Salagean, Janos Vörös, Stéphanie

Pasche, Susan M. De Paul, Hugh A. Spikes, Marcus Textor and Nicholas D. Spencer,

“Boundary Lubrication of Oxide Surfaces by Poly(L-lysine)-g-poly(ethylene glycol)

(PLL-g-PEG) in Aqueous Media”, Tribology Letters, 2003, 10.1023/A:1024861119372

4. Markus Müller, Seunghwan Lee, Hugh A. Spikes and Nicholas D. Spencer, “The

Influence of Molecular Architecture on the Macroscopic Lubrication Properties of the

Brush-Like Co-polyelectrolyte Poly(L-lysine)-g-poly(ethylene glycol) (PLL-g-PEG)

Adsorbed on Oxide Surfaces”, Tribology Letters, 2003,

10.1023/B:TRIL.0000003063.98583.bb

5. Xiaoping Yan, Scott S. Perry, Nicholas D. Spencer, Stéphanie Pasche, Susan M. De

Paul, Marcus Textor and Min Soo Lim, “Reduction of Friction at Oxide Interfaces upon

Polymer Adsorption from Aqueous Solutions”, Langmuir, 2004, 10.1021/la035785b

6. Seunghwan Lee, Rico Iten, Markus Müller and Nicholas D. Spencer, “Influence of

Molecular Architecture on the Adsorption of Poly(ethylene oxide)-Poly(propylene

oxide)-Poly(ethylene oxide) on PDMS Surfaces and Implications for Aqueous

Lubrication”, Macromolecules, 2004, 10.1021/ma049076w

7. Motoyasu Kobayashi, Yuki Terayama, Nao Hosaka, Masataka Kaido, Atsushi Suzuki,

Norifumi Yamada, Naoya Torikai, Kazuhiko Ishihara and Atsushi Takahara, “Friction

behavior of high-density poly(2-methacryloyloxyethyl phosphorylcholine) brush in

Page 46: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

45

aqueous media”, Soft Matter, 2007, 10.1039/b615780g

8. Meng Chen, Wuge H. Briscoe, Steven P. Armes and Jacob Klein, “Lubrication at

Physiological Pressures by Polyzwitterionic Brushes”, Science, 2009,

10.1126/science.1169399

9. Chiara Perrino, Seunghwan Lee, Sung Won Choi, Atsushi Maruyama and Nicholas D.

Spencer, “A Biomimetic Alternative to PEG as an Antifouling Coating: Resistance to

Non-Specific Protein Adsorption of Poly(L-lysine)-graft-Dextran”, Langmuir, 2008,

10.1021/la800947z

10. Chiara Perrino, Seunghwan Lee and Nicholas D. Spencer, “Surface-grafted Sugar Chains

as Aqueous Lubricant Additives: Synthesis and Macrotribological Tests of Poly(L-

lysine)-graft-Dextran (PLL-g-dex) Copolymers”, Tribology Letters, 2009,

10.1007/s11249-008-9402-6

11. Stephen P. Massia and John Stark, “Immobilized RGD peptides on surface-grafted

dextran promote biospecific cell attachment”, Journal of Biomedical Materials Research,

2001, 10.1002/1097-4636(20010905)56:3<390::AID-JBM1108>3.0.CO;2-L

12. Surangkhana Martwiset, Anna E. Koh and Wei Chen, “Nonfouling Characteristics of

Dextran-Containing Surfaces”, Langmuir, 2006, 10.1021/la061064b

13. Jacob Piehler, Andreas Brecht, Karin Hehl and Günter Gauglitz, “Protein interactions in

covalently attached dextran layers”, Colloids and Surfaces B: Biointerfaces, 1999,

10.1016/S0927-7765(99)00046-6

14. Nolan B. Holland, Yongxing Qiu, Mark Ruegsegger and Roger E. Marchant,

“Biomimetic engineering of non-adhesive glycocalyx-like surfaces using oligosaccharide

surfactant polymers”, Nature, 1998, 10.1038/33894

15. R. A. Frazier, G. Matthijs, M. C. Davies, C. J. Roberts, E. Schacht and S. J. B. Tendler,

“Characterization of protein-resistant dextran monolayers”, Biomaterials, 2000,

Page 47: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

46

10.1016/S0142-9612(99)00270-7

16. Anwarul Ferdous, Hiromitsu Watanabe, Toshihiro Akaike and Atsushi Maruyama,

Poly(L-lysine)–graft–dextran copolymer: amazing effects on triplex stabilization under

physiological pH and ionic conditions (in vitro), Nucleic Acids Research, 1998, 26, 3949

17. Atsushi Maruyama, Hiromitsu Watanabe, Anwarul Ferdous, Maiko Katoh, Tsutomu

Ishihara and Toshihiro Akaike, “Characterization of interpolyelectrolyte complexes

between double-stranded DNA and polylysine comb-type copolymers having hydrophilic

side chains”, Bioconjugate Chemistry, 1998, 10.1021/bc9701510

18. Seunghwan Lee and Nicholas D. Spencer, “Sweet, Hairy, Soft, and Slippery”, Science,

2008, 10.1126/science.1153273

19. Scott S. Perry, Xiaoping Yan, F. T. Limpoco, Seunghwan Lee, Markus Müller and

Nicholas D. Spencer, “Tribological Properties of Poly(L-lysine)-graft-poly(ethylene

glycol) Films: Influence of Polymer Architecture and Adsorbed Conformation”, ACS

Applied Materials and Interfaces, 2009, 10.1021/am900101m

20. Gary S. Grest, “Interfacial Sliding of Polymer Brushes: A Molecular Dynamics

Simulation”, Physical Review Letters, 1996, 10.1103/PhysRevLett.76.4979

21. Florent Goujon, Patrice Malfreyt and Dominic J. Tildesley, “Mesoscopic simulation of

entangled polymer brushes under shear: compression and rheological properties”,

Macromolecules, 2009, 10.1021/ma9000429

22. Manoj K. Chaudhury and George M. Whitesides, “How to Make Water Run Uphill”,

Science, 1992, 10.1126/science.256.5063.1539

23. Jan Genzer, “Templating surfaces with gradient assemblies”, Journal of Adhesion, 2005,

10.1080/00218460590944855

24. Nripen Singh, Xiaofeng Cui, Thomas Boland and Scott M. Husson, “The role of

independently variable grafting density and layer thickness of polymer nanolayers on

Page 48: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

47

peptide adsorption and cell adhesion”, Biomaterials, 2007,

10.1016/j.biomaterials.2006.09.036

25. Tao Wu, Kirill Efimenko, Petr Vlcek, Vladimír Šubr and Jan Genzer, “Formation and

Properties of Anchored Polymers with a Gradual Variation of Grafting Densities on Flat

Substrates”, Macromolecules, 2003, 10.1021/ma0257189

26. C. F. Lane, “Sodium Cyanoborohydride - A Highly Effective Reducing Agent for

Organic Functional Compounds”, Synthesis, 1975, 135

27. Richard F. Borch, Mark D. Bernstein and H. Dupont Durst, “Cyanohydridoborate anion

as a selective reducing agent”, Journal of the American Chemical Society, 1971,

10.1021/ja00741a013

28. Sara Morgenthaler, Christian Zink, Brigitte Städler, Janos Vörös, Seunghwan Lee,

Nicholas D. Spencer and Samuele G. P. Tosatti, “Poly(L-lysine)-grafted-poly(ethylene

glycol)-based surface-chemical gradients. Preparation, characterization, and first

applications”, Biointerphases, 2006, 10.1116/1.2431704

29. Lydia M. Feller, Simona Cerritelli, Marcus Textor, Jeffrey A. Hubbell and Samuele G. P.

Tosatti, “Influence of Poly(propylene sulfide-block-ethylene glycol) Di- and Triblock

Copolymer Architecture on the Formation of Molecular Adlayers on Gold Surfaces and

Their Effect on Protein Resistance: A Candidate for Surface Modification in Biosensor

Research”, Macromolecules, 2005, 10.1021/ma051424m

30. Fredrik Höök, Janos Vörös, M. Rodahl, R. Kurrat, P. Böni, J. J. Ramsden, Markus

Textor, Nicholas D. Spencer, P. Tengvall, J. Gold and B. Kasemo, “A comparative study

of protein adsorption on titanium oxide surfaces using in situ ellipsometry, optical

waveguide lightmode spectroscopy, and quartz crystal microbalance/dissipation”,

Colloids and Surfaces B : Biointerfaces, 2002, 10.1016/S0927-7765(01)00236-3

31. Janos Vörös, J. J. Ramsden, G. Csucs, I. Szendro, Markus Textor and Nicholas D.

Spencer, “Optical grating coupler biosensors”, Biomaterials, 2002, 10.1016/S0142-

Page 49: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

48

9612(02)00103-5

32. R.Kurrat, Markus Textor, J. J. Ramsden, P. Boni and Nicholas D. Spencer, “Instrumental

improvements in optical waveguide light mode spectroscopy for the study of

biomolecule adsorption”, Review of Scientific Instruments, 1997, 10.1063/1.1148069

33. Juai Ruan and Bharat Bhushan, “Nanoindentation studies of sublimed fullerene films

using atomic force microscopy”, Journal of Materials Research, 1993, 8, 3019-22

34. S. M. Cook, T. E. Schäffer, K. M. Chynoweth, M. Wigton, R. W. Simmonds and

Kristine M. Lang, “Practical implementation of dynamic methods for measuring atomic

force microscope cantilever spring constants”, Nanotechnology, 2006, 10.1088/0957-

4484/17/9/010

35. John E. Sader, James W. M. Chon and Paul Mulvaney, “Calibration of rectangular

atomic force microscope cantilevers”, Review of Scientific Instruments, 1999,

10.1063/1.1150021

36. Rachel J. Cannara, Michael Eglin and Robert W. Carpick, “Lateral force calibration in

atomic force microscopy: A new lateral force calibration method and general guidelines

for optimization”, Review of Scientific Instruments, 2006, 10.1063/1.2198768

37. John E. Sader and Christopher P. Green, “In-plane deformation of cantilever plates with

applications to lateral force microscopy”, Review of Scientific Instruments, 2004,

10.1063/1.1667252

38. Malvern Zetasizer Nano application note MRK839-01, 2007

39. Sabine M. Görisch, Malte Wachsmuth, Katalin Fejes Tóth, Peter Lichter and Karsten

Rippe, “Histone acetylation increases chromatin accessibility”, Journal of Cell Science,

2005, 10.1242/jcs.02689

40. Tao Wu, Peng Gong, Igal Szleifer, Petr Vlcek, Vladimír Šubr and Jan Genzer, “Behavior

of Surface-Anchored Poly(acrylic acid) Brushes with Grafting Density Gradients on

Page 50: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

49

Solid Substrates: 1. Experiment”, Macromolecules, 2007, 10.1021/ma0710176

41. Ivan U. Vakarelski, Scott C. Brown, Yakov I. Rabinovich and Brij M. Moudgil, “Lateral

Force Microscopy Investigation of Surfactant-Mediated Lubrication from Aqueous

Solution”, Langmuir, 2004, 10.1021/la0352873

42. Torbjörn Pettersson, Ali Naderi, Ricardas Makuška and Per M. Claesson, “Lubrication

Properties of Bottle-Brush Polyelectrolytes: An AFM Study on the Effect of Side Chain

and Charge Density”, Langmuir, 2008, 10.1021/la703229n

43. Shinpei Yamamoto, Muhammad Ejaz, Yoshinobu Tsujii and Takeshi Fukuda, “Surface

Interaction Forces of Well-Defined, High-Density Polymer Brushes Studied by Atomic

Force Microscopy. 2. Effect of Graft Density”, Macromolecules, 2000,

10.1021/ma991988o

44. Yuquan Zou, Nicholas A. A. Rossi, Jayachandran N. Kizhakkedathu and Donald E.

Brooks, “Barrier Capacity of Hydrophilic Polymer Brushes To Prevent Hydrophobic

Interactions: Effect of Graft Density and Hydrophilicity”, Macromolecules, 2009,

10.1021/ma901055t

45. Jacob Klein, Eugenia Kumacheva, Diana Mahalu, Dvora Perahia and Lewis J. Fetters,

“Reduction of frictional forces between solid surfaces bearing polymer brushes”, Nature,

1994, 10.1038/370634a0

46. Nir Kampf, Jean-François Gohy, Robert Jérôme and Jacob Klein, “Normal and shear

forces between a polyelectrolyte brush and a solid surface”, Journal of Polymer Science

Part B: Polymer Physics, 2005, 10.1002/polb.20321

47. Uri Raviv, Suzanne Giasson, Nir Kampf, Jean-François Gohy, Robert Jérôme and Jacob

Klein, “Lubrication by charged polymers”, Nature, 2003, 10.1038/nature01970

48. Benoît Liberelle and Suzanne Giasson, “Friction and Normal Interaction Forces between

Irreversibly Attached Weakly Charged Polymer Brushes”, Langmuir, 2008,

Page 51: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

50

10.1021/la702367f

49. Marcel Benz, Nianhuan Chen, Gregory Jay and Jacob Israelachvili, “Static Forces,

Structure and Flow Properties of Complex Fluids in Highly Confined Geometries”,

Annals of Biomedical Engineering, 2005, 10.1007/s10439-005-8961-z

50. Kenneth J. Rosenberg, Tolga Goren, Rowena Crockett and Nicholas D. Spencer, “Load-

induced transitions in the lubricity of adsorbed poly(L-lysine)-g-dextran as a function of

polysaccharide chain density”, ACS Applied Materials and Interfaces, 2011,

10.1021/am200521m

Page 52: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

51

Chapter 4: Influence of solutes on

hydration and lubricity of dextran brushes

This work was published in 2012 in CHIMIA.28

The work described was done by myself,

under the supervision of Dr. Rowena Crockett and Prof. Nicholas D. Spencer, who also

contributed in discussions about the project steps and corrected the manuscript.

4.1 Abstract

The characteristic lubricity and non-fouling behavior of polymer brushes is critically

dependent on the solvation of the polymer chains, as well as the chain-chain interactions.

Dextran brushes have shown promise as non-toxic aqueous lubricant films, and are similar in

composition to natural lubricating systems, while their comparative simplicity allows for

controlled preparation and fine characterization. This project entails measuring the solvation

and lubricity of dextran brushes in the presence of additives which modify the inter-chain

hydrogen bonding. The thickness and refractive index of the film were measured during

adsorption of the brush layer onto a silica substrate and the subsequent immersion in

solutions of potassium sulfate and α,α-trehalose. We also studied the lubricity of the system

as a function of normal loading using colloidal-probe AFM. Both solutes are shown to have a

minimal effect on the hydration of the brush while significantly reducing the brush lubricity,

indicating that inter-chain hydrogen bonding supports the load-bearing capacity of

polysaccharide brushes.

4.2 Introduction

Polymer brushes have shown great promise as lubricious surface coatings for aqueous

tribosystems.1-3

Composed of densely packed end-attached polymer chains, their high degree

of solvation and load-bearing capacity act to reduce friction and wear, even in boundary-

lubrication conditions, where the hydrodynamic fluid film alone would be insufficient to

maintain separation between opposing asperities. When the brush is comprised of non-toxic

polymers, such as PEG or dextran, the resulting system has great potential in the fields of

food and drug processing, biomedical engineering, and microfluidics.4

Page 53: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

52

The study of polymer brush-mediated aqueous tribosystems is also interesting because they

share some common properties with naturally occurring lubricious systems, such as the

glycocalyx,5 articular cartilage joints,

6 and slippery surfaces such as the tongue,

7 which are

comprised of densely arranged oligosaccharide-bearing biomolecule-coated surfaces

separated by aqueous lubricants such as blood plasma, synovial fluid, or mucus. The

complexity and sensitivity of these systems render the study of their mechanisms of

lubrication highly challenging. Polymer- and especially polysaccharide-based brushes

represent simpler, homogeneous systems that capture some of the critical aspects of natural

lubrication.

Dextran is a naturally occurring glucan with high water solubility. The orientation and

degrees of freedom conferred by its (16) linkage hinder the formation of stable lamellar

structures by the relatively flexible chains, making it suitable for the formation of brushes.

Poly(L-lysine)-graft-dextran (PLL-g-dex)8,9

is a bottlebrush copolymer formed by grafting

some of the primary amine groups of a poly(L-lysine) backbone to the terminal aldehydes of

dextran chains via reductive amination. The unreacted amines maintain a positive charge in

aqueous solution at neutral pH, and the backbone will spontaneously adsorb onto negatively

charged surfaces at multiple locations, while the uncharged, hydrophilic dextran chains

extend into solution. On silicon oxide surfaces in water, these molecules adsorb

spontaneously with sufficient density that the dextran inter-chain spacing is significantly

lower than its radius of gyration, forcing the chains into a brush conformation.

Dextran brushes have been shown to have non-fouling properties and to be highly lubricious

at low loads.9,10

Above a critical load, however, the friction increases sharply.10

This critical

load is attributed to the point at which the adjacent chains of the brush are forced apart,

disrupting the hydrogen bonds, which may then stick to the counterface. Tribosystems

composed of opposing dextran brushes show lower friction at low loads, a higher critical

load, and higher friction above the critical load than dextran brushes against bare silica.

Similar behavior can be seen in models of brush-brush interaction where the chains are long

and flexible enough to entangle.11-14

This is not possible for the shorter, stiffer dextran chains

studied, but the inter-chain hydrogen bonds are hypothesized to cause this similar behavior.

In order to better understand the role of inter-chain hydrogen bonding in polysaccharide

brushes, we measured the lubricity and swelling of the brush in the presence of solutes that

influence hydrogen bonding. Potassium sulfate is an ionic kosmotrope that can screen

Page 54: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

53

electrostatic interactions and reduce hydrogen bonding by enhancing dextran’s affinity to

water.15,16

Mono- and disaccharides, such as glucose and α,α-trehalose, can also reduce inter-

chain cohesion by acting as "zippers", as they bridge dextran chains and their mobility allows

increased freedom to restructure the hydrogen bonds.17,18

These and similar low-molecular-

weight saccharides and polyols are often used as a plasticizers in polysaccharide melts.

Trehalose is extremely hydrophilic, with a hydration shell four times as large of that of

glucose, and a strong kosmotrope, making it a powerful bioprotectant against plant

dehydration.19

The friction, and especially the transition load, of the dextran brush in the

presence of these solutes should be a strong indicator of their effects on the inter-chain

hydrogen bonding.

4.3 Experimental

4.3.1 PLL-g-dextran brushes

The PLL-g-dex copolymer synthesis via reductive amination has been described

previously.8,9

Dextran (dextran T5, 5.2 kDa, polydispersity 1.8, Pharmacosmos A/S,

Denmark) was mixed with poly(L-lysine)-HBr (20 kDa, polydispersity 1.1, Sigma-Aldrich,

Switzerland) in a sodium borate buffer (0.1 M, pH 8.5, 0.4 M NaCl). The terminal aldehydes

of the dextran chains formed a Schiff base with the primary amines of the PLL, which was

then reduced with sodium cyanoborohydride. The product was separated from the reactants

by filtration with ultracentrifugation. The grafting ratio of lysine monomers to dextran chains

was determined to be 5.3 via 1H-NMR in D2O and elemental analysis.

4.3.2 Optical measurements of brush thickness and refractive

index

Optical waveguide lightmode spectrometry (OWLS, Microvacuum, Hungary) was used to

monitor the adsorption of the brush and subsequent exposure to trehalose solution in real

time. The standard Si0.75Ti0.25O2 chip was ultrasonicated in COBAS (Roche, Switzerland) for

10 min, rinsed in ultrapure water, ultrasonicated in 2-propanol, dried in a nitrogen stream, and

cleaned in a UV/ozone cleaner (Bioforce Nanosciences, Iowa, USA) for 30 min. The cleaned

waveguide was incubated in ultrapure water in the flow cell overnight to achieve a stable

baseline. The solution of 0.02 g/L PLL-g-dex in ultrapure water was then introduced, and the

Page 55: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

54

sample allowed to stabilize until the change was less than the instrument experimental drift,

then rinsed with ultrapure water, allowed to stabilize again, and finally exposed to the 5.5

mM trehalose solution and allowed to stabilize. The thickness and refractive index given by

the instrument were used to calculate the optical mass with the de Feijter equation,23

assuming a refractive index increment dn/dc of 0.131 cm3/g.

8,9 As the experiment is

performed in aqueous conditions, the water associated with the polymer layer is not included

in the measured optical mass, referred to as the “dry mass”.

Transmission interference adsorption spectroscopy (TInAS)20,21

was used to monitor the

adsorption of the brush and subsequent exposure to K2SO4 solution in real time. A standard

TInAS sensor, consisting of glass coated with 25 nm Al and 3 μm SiO2, was ultrasonicated

twice for 5 minutes in toluene, then twice for 5 minutes in 2-propanol, then dried with

nitrogen and exposed to oxygen plasma (Plasma Cleaner/Sterilizer PDC-32G, Harrick, NY,

USA) for 2 min. The cleaned sensor was incubated in ultrapure water overnight to achieve a

stable baseline. Then, the solution of 0.02 g/L PLL-g-dex in ultrapure water was introduced,

and the sample allowed to stabilize until the change was less than the instrument

experimental drift, then rinsed with ultrapure water, allowed to stabilize again, and finally

exposed to the 10 mM K2SO4 solution and allowed to stabilize. The dry optical mass was

calculated from the obtained thickness and refractive index in the same way as with the

OWLS.

Variable angle spectroscopic ellipsometry (VASE, J. A. Woolam Co, Inc., NE, USA) was

used to check the adsorption of the brush onto the silicon wafers used for the friction

measurements. Bare wafers were cleaned by the same procedure as the TInAS sensors, then

rinsed in ultrapure water and dried with nitrogen. The thickness of the native oxide was

measured in multiple places at wavelengths 370-1000 nm at an incident angle of 70°. The

thickness variation over the 1 cm2 wafers was less than 0.5 Å. The substrates were then

exposed for different lengths of time to the PLL-g-dex solutions, rinsed in ultrapure water,

and dried with nitrogen. The dextran brush thickness was then measured in several places

under the same conditions, and also showed variation over the wafers of less than 0.5 Å. In

contrast to OWLS and TInAS, the VASE measurement is performed in air, rather than in

aqueous solution, and therefore also includes any water associated with the film.

Additionally, the effective refractive index of the film is not measured by the VASE and must

be assumed. For these reasons, the measured thickness does not agree quantitatively with

TInAS/OWLS measurements, and the adsorbed dry mass cannot be directly calculated.

Page 56: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

55

However, in previous work,10

the VASE thickness (assuming an effective refractive index of

1.45) was found to scale linearly with the TInAS/OWLS dry mass. Therefore the VASE

thickness was multiplied by the scaling proportion to determine the dry optical mass on the

wafers, and thereby the L/2Rg values, of the dextran layers, with resulting values of L/2Rg

ranging from 1.04 to 0.5.

4.3.3 Colloidal-probe lateral force measurements

The friction measurements were conducted using an MFP-3D AFM (Asylum Research,

California, USA). Borosilicate colloidal spheres (Kromasil, Eka Chemicals AB, Sweden;

radius ≈ 8 μm) were glued to tipless rectangular cantilevers (CSC12, Mikromasch, Estonia)

with UV-curable epoxy (Norland Optical Adhesive 63, Norland, NJ, USA). The glue was

cured for 15 minutes under UV and then allowed to set overnight at 50°C. Prior to attaching

the sphere, the normal and lateral spring constants of the sphere were calibrated from the

power spectral density of the thermal noise fluctuations24

and the method of Sader et al.25

using the measured cantilever dimensions.

Each bare colloid was first slid against a bare substrate to ensure that no transient effects,

such as initial colloidal wear or initial laser drift, could influence the measurements. Friction

measurements were made in contact mode by sliding the probe perpendicular to the

cantilever axis over a distance of 1 μm at a speed of 1 μm/s. The trace minus retrace (TMR)

signal was recorded, which is directly proportional to the friction force.26

Fifteen curves were

taken for each load, after which the position on the sample was shifted slightly to avoid any

sample wear and the load increased for the next step. To obtain the friction force, the TMR

signal was multiplied by the calibration constant determined using the test probe method of

Carpick et al.27

The measurements shown in Figure 4.1 were made in a single session without

disturbing the cantilever or laser, and likewise for Figure 4.2, so the calibration does not

affect their relative values, but only facilitates the comparison between the two datasets,

shown in Figure 4.3, and with other studies.

4.4 Results and discussion

4.4.1 Brush solvation

The sensors of two different optical instruments, the optical waveguide lightmode

spectroscope (OWLS) and the transmission interference adsorption sensor (TInAS),20,21

were

Page 57: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

56

exposed to a solution of PLL-g-dex(5 kDa) in ultrapure water, and the thickness and

refractive index of the film were measured in real time during adsorption as the brush was

formed. The adsorbed mass of the polymer on the surface, not including any associated water

(the “dry mass”), were calculated from the thickness and refractive index. The two

instruments generally show excellent agreement, in this case reporting the maximum

adsorbed mass of 306 ± 15 ng/cm2, corresponding to an average inter-chain distance L of 1.9

nm or a grafting density L/2Rg of 0.5, taking the radius of gyration Rg to also be 1.9 nm.22

The measurements indicate a mean solvated brush thickness of 7.1 nm with a refractive index

of 1.42, although these values have a higher mutual uncertainty than their combined optical

mass.21

After adsorption had slowed down to within the drift of the instruments, the samples were

first rinsed with ultrapure water and then exposed to a solution of either 10 mM potassium

sulfate or 5.5 mM trehalose in ultrapure water. The brush showed no measurable change in

thickness or refractive index in the presence of either solute, indicating that the solvent

quality did not significantly decrease, which would cause the brush to collapse, nor did the

solute cause any measurable desorption of the copolymer.

4.4.2 Aqueous lubricity

Silicon wafers were incubated in the same PLL-g-dex solution for varying times to achieve

different chain densities. They were subsequently rinsed and dried, and the thickness of the

film measured with variable angle spectroscopic ellipsometry (VASE) to determine the

adsorbed mass and the homogeneity of the films. Finally, the friction force was measured as a

function of load by means of lateral force microscopy, between the brushes and a borosilicate

colloidal sphere glued to the cantilever tip. Measurements were performed in ultrapure water

and subsequently in the appropriate solution.

4.4.3 Influence of grafting density

Figure 4.1 shows the friction response as a function of applied load of PLL-g-dex films of

different densities. In good agreement with an earlier study,10

the local coefficient of friction,

μ = dFlat/dFnorm, is lowest at low loads, but increases sharply above a critical load. Higher

dextran grafting density results in lower friction at low loads and a higher critical load. This

critical load is the value at which the applied load partially overcomes the inter-chain bonds

and steric forces that force the chains into a brush configuration. Above this load, phenomena

Page 58: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

57

such as plowing and dragging overcome the tendency of the brush to maintain a lubricious

aqueous film at the surface, resulting in sharply increased friction.

Figure 4.1 Friction force measurements of a colloidal probe sliding on a silicon wafer coated

with PLL-g-dextran at a grafting density of L/2Rg = 1.04 (dotted line), L/2Rg = 0.68 (dashed

line), and L/2Rg = 0.5 (solid lines), in ultrapure water ( ) and in 5.5 mM trehalose ( ).

4.4.3 Influence of solutes

The friction response of the densest dextran brush in the presence of 5.5 mM trehalose

solution was similar to that in ultrapure water at the lowest loads, but continuously increased

with increasing load, exhibiting a less well-defined critical load (Figure 4.1). At the highest

normal force measured in the presence of trehalose, the friction force was around double that

in ultrapure water. This is consistent with our prediction that the trehalose acts as a plasticizer

for the dextran chains, easing their separation and thus causing a reduction in load-bearing

capacity.

The influence of potassium sulfate on the friction of the densest dextran brush is similar to

that of trehalose, as shown in Figures 4.2 and 4.3. The friction force and local friction

coefficient are similar to the response in ultrapure water at the lowest applied loads, but

0

50

100

150

200

250

300

350

400

450

0 50 100 150 200 250 300 350

Late

ral f

orc

e, n

N

Normal force, nN

Page 59: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

58

diverge at higher loads. A similar trend is already apparent in Figure 4.2, which shows the

local coefficient of friction diverging more rapidly from the initial value with increasing salt

concentrations. Figure 4.3 also shows that the dense brush in the presence of solutes exhibits

a higher critical load than a sparser brush in water with similar friction values. This difference

may reflect the contribution to the load-bearing capacity of the brush made by steric forces,

which are only weakly affected by the presence of solutes.

Figure 4.2 Friction force measurements of a colloidal probe sliding on a silicon wafer coated

with PLL-g-dextran at a grafting density of L/2Rg = 0.5 in ultrapure water ( ), 2 mM K2SO4

( ), and 10 mM K2SO4 ( ).

0

20

40

60

80

100

0 50 100 150 200 250 300 350

Late

ral f

orc

e, n

N

Normal force, nN

Page 60: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

59

Figure 4.3 Data from Figures 4.1 and 4.2 overlaid for comparison. The friction response of

the dense dextran brush in the presence of solutes (filled shapes) is comparable to the

response of sparser dextran films in ultrapure water (broken lines).

4.5 Conclusions

We have demonstrated that the lubricity of dextran brushes can be influenced by solutes that

reduce the strength of the inter-chain hydrogen bonding, without influencing the height or

density of the brush in the absence of load. In particular, the solutes reduce the ability of the

dextran brush to support high loads, while having less effect on the friction at low loads than

a corresponding reduction in brush density, illustrating the relative contributions of inter-

chain hydrogen bonding and steric forces to the lubricity of the brush as a function of applied

load. This has implications for the model of physiological lubrication by saccharide-bearing,

brush-like systems, which usually operate in the presence of much higher salt concentrations.

In biological systems, the presence of other species that enhance the hydrogen bonding or

directly cross-link the chains may serve to offset the failure of inter-chain hydrogen bonding

under those conditions. Similar techniques might be used to extend the lubricious regime of

0

50

100

150

200

250

300

350

400

450

0 50 100 150 200 250 300 350

Late

ral f

orc

e, n

N

Normal force, nN

Page 61: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

60

the dextran brush to higher loads, or to tailor the inter-chain interactions for further

applications such as trapping and then releasing biomolecules.

4.6 References

1. Seunghwan Lee and Nicholas D. Spencer, Chapter 21 in Superlubricity, Elsevier Science

B.V., Amsterdam, 2007

2. Seunghwan Lee, Markus Müller, Monica Ratoi-Salagean, Janos Vörös, Stéphanie

Pasche, Susan M. De Paul, Hugh A. Spikes, Marcus Textor and Nicholas D. Spencer,

“Boundary Lubrication of Oxide Surfaces by Poly(L-lysine)-g-poly(ethylene glycol)

(PLL-g-PEG) in Aqueous Media”, Tribology Letters, 2003, 10.1023/A:1024861119372

3. Wuge H. Briscoe, Simon Titmuss, Fredrik Tiberg, Robert K. Thomas, Duncan J.

McGillivray and Jacob Klein, “Boundary lubrication under water”, Nature, 2006,

10.1038/nature05196

4. Duncan Dowson, “History of Tribology”, Professional Engineering Publishing, London,

1998.

5. Qian Yang, Christian Kaul and Mathias Ulbricht, “Anti-nonspecific Protein Adsorption

Properties of Biomimetic Glycocalyx-like Glycopolymer Layers: Effects of

Glycopolymer Chain Density and Protein Size”, Langmuir, 2010, 10.1021/la903895q

6. Marcella Roba, Marco Naka, Emanuel Gautier, Nicholas D. Spencer and Rowena

Crockett, “The adsorption and lubrication behavior of synovial fluid proteins and

glycoproteins on the bearing-surface materials of hip replacements”, Biomaterials, 2009,

10.1016/j.biomaterials.2008.12.062

7. Seunghwan Lee and Nicholas D. Spencer, “Sweet, Hairy, Soft, and Slippery”, Science,

2008, 10.1126/science.1153273

8. Chiara Perrino, Seunghwan Lee, Sung Won Choi, Atsushi Maruyama and Nicholas D.

Spencer, “A Biomimetic Alternative to PEG as an Antifouling Coating: Resistance to

Page 62: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

61

Non-Specific Protein Adsorption of Poly(L-lysine)-graft-Dextran”, Langmuir, 2008,

10.1021/la800947z

9. Chiara Perrino, Seunghwan Lee and Nicholas D. Spencer, “Surface-grafted Sugar Chains

as Aqueous Lubricant Additives: Synthesis and Macrotribological Tests of Poly(L-

lysine)-graft-Dextran (PLL-g-dex) Copolymers”, Tribology Letters, 2009,

10.1007/s11249-008-9402-6

10. Kenneth J. Rosenberg, Tolga Goren, Rowena Crockett and Nicholas D. Spencer, “Load-

induced transitions in the lubricity of adsorbed poly(L-lysine)-g-dextran as a function of

polysaccharide chain density”, ACS Applied Materials and Interfaces, 2011,

10.1021/am200521m

11. Jacob Klein, in “Fundamentals of Tribology and Bridging the Gap Between the Macro-

and Micro/Nanoscales”, Kluwer Academic Pub., Dordrecht, 2001, p. 177-198

12. Uri Raviv, Suzanne Giasson, Nir Kampf, Jean-François Gohy, Robert Jérôme and Jacob

Klein, “Lubrication by charged polymers”, Nature, 2003, 10.1038/nature01970

13. Benoît Liberelle and Suzanne Giasson, “Friction and Normal Interaction Forces between

Irreversibly Attached Weakly Charged Polymer Brushes”, Langmuir, 2008,

10.1021/la702367f

14. Florent Goujon, Patrice Malfreyt and Dominic J. Tildesley, “Mesoscopic simulation of

entangled polymer brushes under shear: compression and rheological properties”,

Macromolecules, 2009, 10.1021/ma9000429

15. B. Yu. Zaslavsky, T. O. Bagirov, A. A. Borovskaya, N. D. Gulaeva, L. H. Miheeva, A.

U. Mahmudov and M. N. Rodnikova, “Structure of water as a key factor of phase

separation in aqueous mixtures of two nonionic polymers”, Polymer, 1989,

10.1016/0032-3861(89)90301-7

16. B. Yu. Zaslavsky, A. U. Mahmudov, T. O. Bagirov, A. A. Borovskaya, G. Z. Gasanova,

N. D. Gulaeva, V. Yu. Levin, N. M. Mestechkina, L. M. Miheeva and M. N. Rodnikova,

Page 63: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

62

“Aqueous biphasic systems formed by nonionic polymers II. Concentration effects of

inorganic salts on phase separation”, Colloid and Polymer Science, 1987,

10.1007/BF01412510

17. Yachuan Zhang and J. H. Han, “Plasticization of Pea Starch Films with Monosaccharides

and Polyols”, Journal of Food Science, 2006, 10.1111/j.1750-3841.2006.00075.x

18. Yachuan Zhang and J. H. Han, “Crystallization of High-Amylose Starch by the Addition

of Plasticizers at Low and Intermediate Concentrations”, Journal of Food Science, 2010,

10.1111/j.1750-3841.2009.01404.x

19. Matthias Heyden, Erik Bründermann, U. Heugen, Gudrun Niehues, David M. Leitner

and Martina Havenith, “Long-Range Influence of Carbohydrates on the Solvation

Dynamics of Water - Answers from Terahertz Absorption Measurements and Molecular

Modeling Simulations”, Journal of the American Chemical Society, 2008,

10.1021/ja0781083

20. Manfred Heuberger and Tobias E. Balmer, “The transmission interferometric adsorption

sensor”, Journal of Physics D: Applied Physics, 2007, 10.1088/0022-3727/40/23/S07

21. Takumi Sannomiya, Tobias E. Balmer, Manfred Heuberger and Janos Vörös,

“Simultaneous refractive index and thickness measurement with the transmission

interferometric adsorption sensor”, Journal of Physics D: Applied Physics, 2010,

10.1088/0022-3727/43/40/405302

22. Malvern Zetasizer Nano application note MRK839-01, 2007

23. J. A. De Feijter, J. Benjamins and F. A. Veer, “Ellipsometry as a tool to study the

adsorption behavior of synthetic and biopolymers at the air–water interface”,

Biopolymers, 1978, 10.1002/bip.1978.360170711

24. S. M. Cook, T. E. Schäffer, K. M. Chynoweth, M. Wigton, R. W. Simmonds and

Kristine M. Lang, “Practical implementation of dynamic methods for measuring atomic

force microscope cantilever spring constants”, Nanotechnology, 2006, 10.1088/0957-

Page 64: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

63

4484/17/9/010

25. John E. Sader, James W. M. Chon and Paul Mulvaney, “Calibration of rectangular

atomic force microscope cantilevers”, Review of Scientific Instruments, 1999,

10.1063/1.1150021

26. Juai Ruan and Bharat Bhushan, “Nanoindentation studies of sublimed fullerene films

using atomic force microscopy”, Journal of Materials Research, 1993, 8, 3019-22

27. Rachel J. Cannara, Michael Eglin and Robert W. Carpick, “Lateral force calibration in

atomic force microscopy: A new lateral force calibration method and general guidelines

for optimization”, Review of Scientific Instruments, 2006, 10.1063/1.2198768

28. Tolga Goren, Rowena Crockett and Nicholas D. Spencer, “Influence of Solutes on

Hydration and Lubricity of Dextran Brushes”, CHIMIA, 2012, 10.2533/chimia.2012.192

Page 65: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

64

Chapter 5: Impact of chain morphology on

the lubricity of surface-grafted

polysaccharides

This work is from a manuscript in preparation to be submitted to an upcoming special issue

of RSC Advances. The supplementary materials have been added to the end of this chapter.

The work described was done by myself. Dr. Rowena Crockett and Prof. Nicholas D. Spencer

supervised the work, contributed in discussions about the project steps, and corrected the

manuscript.

5.1 Abstract

Surface-density gradients of dextran have been fabricated on silicon wafers by means of an

intermediate azide-terminated monolayer, which binds to random segments of the dextran

chains to form a complex loop-train-tail structure. Friction-force microscopy was employed

to understand the relative contributions of chain density and other parameters on the

tribological behavior of the immobilized chains. The results are contrasted with similar

investigations performed with density gradients of poly(L-lysine)-graft-dextran, a bottlebrush

copolymer that adsorbs onto silica to form a well-characterized dextran brush. Both systems

exhibit friction coefficients that vary over more than an order of magnitude with applied load,

with a sharp transition from low- to high-friction regimes occurring upon increasing load.

The brush architecture exhibited more extreme friction coefficients than the loop-train-tail

architecture, lubricating better at low loads while exhibiting higher friction at high loads,

despite involving less than a third of the amount of dextran (on a monomer basis) in

comparison to the loop-train-tail system. The coefficient of friction at high loads decreased

with increasing dextran surface density in the loop-train-tail system, while the opposite was

true for the polymer brush. The surface density required to forestall the pressure-induced

transition to high friction was also significantly higher for the loop-train-tail system than for

the brush system. These results illustrate the influence of brush regularity on resistance to

collapse under applied load, but also its role in exacerbating friction forces.

Page 66: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

65

5.2 Introduction

Surfaces bearing densely packed, end-grafted neutral or charged polymer chains in good

solvents are extraordinarily resistant to the intrusion of foreign bodies, leading to their

characteristic low-friction and non-fouling behavior.1,2

Particularly interesting are brushes

composed of hydrophilic biocompatible polymers, such as poly(ethylene glycol), which have

been the subject of many investigations as non-fouling surfaces, aqueous lubricants, and so-

called “stealth coatings” against undesired immune-system response.3-6

These characteristics,

observable over a wide range of chain chemistries and architectures, originate from the steric

hindrance presented by the chains and their associated solvent molecules, as well as the

osmotic pressure induced by the strong interactions of the chains with the solvent.7 The

architecture of polymer-brush-mediated aqueous tribosystems also bears a striking similarity

to that of naturally occurring lubricious systems, including the glycocalyx on the inner walls

of capillaries,8 the surface of articular cartilage,

9 and slippery surfaces such as the eye and

tongue.10

Such natural systems typically consist of oligosaccharide-bearing surfaces in

aqueous environments. A characteristic feature of oligo- and polysaccharide chains is the

presence of hydroxyl groups, which can directly engage in hydrogen bonding with adjacent

chains, in contrast to polymers such as poly(ethylene glycol), for which water intercalation is

necessary to bridge between chains. Hydroxyl groups also facilitate selective

functionalization of the chains to produce films with complex architectures by the addition of

reactive groups, branching, or crosslinking. Synthetic polysaccharide brushes and thin films

have been produced by various means, including covalent grafting to silica surfaces,11

Langmuir-Blodgett films of block copolymers,12

and electrostatic adsorption of graft

copolymers.13,14

In order to better understand the roles of the film architecture and inter-chain hydrogen

bonding in polysaccharide films, we compared the friction behavior of two different dextran-

based architectures. Dextran is a naturally occurring hydrophilic polysaccharide consisting of

D-glucose monomers connected by α-1,6 glycosidic linkages, which confer a relatively high

degree of flexibility and prevent the formation of stable crystalline structures. PLL-g-dextran

consists of dextran chains end-grafted to some of the pendant lysine groups of poly(L-

lysine).13-16

The unreacted lysine groups are positively charged under aqueous conditions at

physiological pH, and thus facilitate spontaneous adsorption of the PLL-g-dextran molecule

from solution onto negatively charged surfaces, such as silicon dioxide, with sufficient chain

density to force the hydrophilic dextran chains into a brush conformation, as shown in Figure

Page 67: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

66

5.1. PLL-g-dextran brushes in water are highly lubricious against a hard counterface at low

loads, but the friction increases sharply above a critical load.17,18

When the counterface

consists of a dextran brush, lower friction is obtained at low loads, the critical load is higher,

and friction is higher above the critical load.

Figure 5.1 Chemical structures of poly(L-lysine)-graft-dextran (PLL-g-dextran) (A) and

poly(allylamine)-graft-perfluorophenylazide-dextran (PAAm-g-PFPA-dextran) (B), and the

corresponding expected polymer architectures, when attached to silicon oxide substrates.

A high-coverage, PLL-g-dextran-functionalized surface probably represents the highest

dextran brush density achievable via grafting-to means, with adsorption eventually limited by

steric hindrance of the adsorbed brush, and largely independent of the number of unreacted

Page 68: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

67

lysine “feet” or the dextran or PLL chain lengths.13,14,19

Since there is no readily available

route to achieve higher dextran surface densities in a uniform brush, an alternative method of

attaching dextran to the surface has been pursued. In the present study, substrates were coated

with the adhesion promoter, poly(allylamine)-graft-perfluorophenylazide (PAAm-g-PFPA),

an azide-bearing graft copolymer, and chain-density gradients of dextran chains were

subsequently fabricated, being attached at random points to form a complex loop-train-tail

structure, as shown in Figure 5.1. The influence of the chain density and applied load on the

friction coefficient of the resulting film was investigated, and the results contrasted to the

behavior of PLL-g-dextran brushes.

5.3 Materials and methods

5.3.1 Synthesis of poly(allylamine)-graft-perfluorophenylazide

(PAAm-g-PFPA)

Poly(allylamine)-graft-perfluorophenylazide (PAAm-g-PFPA) was synthesized as previously

described.20

Briefly, poly(allylamine hydrochloride) (6.33 mg, 6.8x10-6

mol monomer, 15

kDa, Sigma) and excess potassium carbonate (15.82 mg, 1.1x10-4

mol, Merck) were

dissolved in 1.3 mL of water (Milli-Q Gradient A10, Millipore SA, Molsheim, France), and a

solution of N-hydroxysuccinimide-perfluorophenylazide (NHS-PFPA, 5.62 mg, 1.7x10-5

mol,

SuSoS AG, Dübendorf, Switzerland) in 1.3 mL of ethanol (Merck) was slowly added. The

resulting solution was stirred overnight and then diluted to 100 mL with a 3:2 (v/v)

ethanol/water mixture, yielding a polymer concentration of 0.1 g/L. All solutions were kept in

the dark to avoid activation of the azide groups. The grafting ratio, defined as the number of

allylamine monomers divided by the number of grafted PFPA monomers, is observed to be 8

or 9.20

5.3.2 Substrate preparation and dextran attachment

Bare 2 cm x 2 cm silicon wafers were cleaned by sonication twice for five minutes in toluene,

then twice for five minutes in 2-propanol, after which they were exposed to O2 plasma for

two minutes (Plasma Cleaner/Sterilizer, PDC-32G instrument, Harrick, Ossining, NY, USA).

The surface SiO2 layer was characterized by ellipsometry, as described below, with none of

the wafers exhibiting more than 0.1 nm variation, after which the wafers were immediately

immersed in a 0.1 g/L solution of PAAm-g-PFPA. After thirty minutes of immersion, the

Page 69: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

68

substrates were rinsed and then sonicated for 5 minutes in a 3:2 (v/v) ethanol/water mixture,

then rinsed with ultrapure water and dried with nitrogen. The PAAm-g-PFPA thickness was

then measured by ellipsometry (described below).

Following this, the wafers were placed on a spin coater, covered in a solution of 25 g/L

dextran (dextran T200, 2 MDa, Pharmacosmos A/S, Denmark) in water, and then spun at

4000 rpm for 40 s, then 5000 rpm for 10 s. The wafers were allowed to dry for 15 minutes,

then exposed to UV light (Philips TUV 11W, peak wavelength 254 nm) at a distance of 10

cm. Dextran gradients were produced by varying the UV exposure time (from 3.5 s to 240 s)

along the wafer using a moving shutter,21

programmed to evenly space the desired exposure

durations in discrete steps along the wafer. Following the UV exposure, the wafers were

immersed overnight in ultrapure water to remove excess non-attached polymer, then rinsed

with water, dried with nitrogen, measured by ellipsometry, and then stored in a dark prior to

friction measurements.

Subsequent UV illumination in ambient air or in air with 100% relative humidity, with the

aim of activating any remaining azides, was found to influence neither the thickness nor the

friction measurements, within experimental error.

5.3.3 Characterization by variable-angle spectroscopic

ellipsometry (VASE)

Thickness measurements of the substrate SiO2 layer, the adsorbed PAAm-g-PFPA layer, and

the attached dextran film were performed in air using an M-2000 variable angle spectroscopic

ellipsometer (VASE, J.A. Woollam Co., Inc., Lincoln, NE, USA) at a reflection angle of 70°

and over the wavelength range of 370-994 nm. The spectra were analyzed using WVASE32

software, using a multilayer model of Si/SiO2 and top films with the optical properties of a

Cauchy layer,22

as detailed in the Supporting Information. The surface density of each layer

can be calculated from the measured thickness and the bulk density of PAAm-g-PFPA and

dextran.20

Ellipsometric measurements on the gradient in water were also made using a liquid

cell, and the results evaluated using the Bruggeman effective-medium approximation,23

whereby the top layer was treated as a mixture of dextran (n=1.51) and water (n=1.333), and

the layer thickness and dextran fraction were both fitted. The multilayer models for the dry

and wet systems are shown in the Supporting Information.

Page 70: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

69

5.3.4 Atomic force microscopy measurements

The AFM experiments were conducted in contact mode by means of a MFP-3D AFM

(Asylum Research, Santa Barbara, CA, USA) with gold-coated rectangular Si cantilevers

(CSC12B, nominal normal spring constant 1.75 N/m, Mikromasch, Tallinn, Estonia), chosen

to match the loading range of interest, while maintaining sensitivity and minimizing errors

due to in-plane bending.24

A borosilicate colloidal sphere (Kromasil, Eka Chemicals AB,

Bohus, Sweden; radius ≈ 8 μm) was glued to the end of the cantilever with UV-curable

epoxy. (Norland Optical Adhesive #63, Norland, Cranbury, NJ, USA) The entire assembly

was then irradiated with UV light for 30 minutes and left for a full day prior to use, to further

strengthen the adhesion between the colloidal sphere and the cantilever.

Friction measurements were obtained by scanning the probe perpendicularly to the major axis

of the cantilever and recording the TMR (Trace Minus Retrace in mV) value. The TMR value

is directly proportional to the friction force and minimizes the contributions to the lateral

force from non-friction sources.25

The gradient was characterized by measuring points in a

random order, in order to minimize possible effects arising from gradual drift. All scans were

1 µm in length with velocities of 1 µm/sec. The dependence of lateral force on scanning

distance was investigated and it was shown that for scans longer than 100 nm, the lateral

force remained constant.

The normal and lateral cantilever spring constants were calibrated prior to sphere attachment

from the power spectral density of the thermal noise fluctuations26

and by the method of

Sader and coworkers.27

The lateral force measurements of the rectangular cantilevers were

calibrated using the test-probe method described by Carpick and coworkers,28

yielding

absolute values for both the lateral force and the local coefficient of friction (the ratio of

lateral to normal force). The test-probe calibration method itself includes correcting factors

for differences in cantilever dimensions, colloidal sphere size, normal and lateral spring

constant of the cantilever, and spot intensity between the experimental and reference setups.

All experiments were measured within a gradient in a single run with the same tip and

cantilever. Additionally, all the rectangular cantilevers used for the calibration and

measurement were produced in a single batch from the same wafer, so that any errors

stemming from non-ideal behavior of the cantilevers or unaccounted factors, such as

variations in the reflectivity of the cantilever top coatings, would be consistent across all

measurements.

Page 71: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

70

Following nanotribological measurements with an uncoated borosilicate colloidal probe, the

entire tip-cantilever assembly was gently lifted off the gradient surface and immediately

exposed to a 0.02 mg/mL PLL-g-dextran (synthesized from previous work)13,14

solution to

form a dextran brush layer on the probe surface. After 30 minutes, the assembly was

thoroughly rinsed with water and placed back into contact with the same gradient substrate to

resume the friction experiments.

5.4 Results and discussion

5.4.1 Gradient fabrication

The VASE dry thickness of the PAAm-g-PFPA layer was 1.7 + 0.1 nm, and the additional

dry thickness of the attached dextran film reached 6 nm at the dense end of the gradient,

corresponding to a density of 36 dextran monomers per nm2, or 343 nm

2 per 2MDa dextran

chain. The kinetics of the UV-activated attachment of dextran to the substrate are seen to

approximate a logarithmic curve until reaching saturation after about one minute, as shown in

Figure 5.2. The aqueous VASE measurements indicate a relatively constant (water + dextran)

film thickness of 124 + 5.4 nm, independent of the UV irradiation time, while the fraction of

dextran within the film increased with increasing UV irradiation, as shown in Figure 5.3.

Films produced with the shorter irradiation times could not be reproducibly measured in the

wet state due to weak signal.

Page 72: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

71

Figure 5.2 Surface density of dextran in PAAm-g-PFPA-dextran films, calculated from

VASE thickness measurements, plotted against UV exposure time.

0

10

20

30

40

1 10 100

Dex

tran

mo

no

me

rs p

er

nm

2

UV exposure, seconds

Page 73: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

72

0

20

40

60

80

100

120

140

160

180

0 10 20 30 40

Laye

r th

ickn

ess

, nm

Dextran monomers per nm2

(A)

0

2

4

6

8

0 10 20 30 40

Laye

r p

erc

en

t d

extr

an

Dextran monomers per nm2

(B)

Page 74: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

73

Figure 5.3 Thickness (A) and mass fraction (B) of the dextran/water layer associated with the

PFPA-coated substrate, measured by VASE in a water-filled liquid cell and calculated using

the Bruggeman mean-field approximation, plotted against the dextran surface density. Mass

fraction is also plotted against UV exposure time (C) for comparison of the kinetics with

those in Figure 5.2.

Comparing the surface densities of dextran and PAAm-g-PFPA calculated from these VASE

measurements at the densest end of the gradient, and taking the experimental grafting ratio of

9 allylamines to each PFPA, we obtain 542 PFPA units per 2 MDa dextran chain, or 56.9

dextran monomers per PFPA unit, with 1.58 PFPA units per nm2. What fraction of these

PFPA units is actually bound to a dextran is unknown. However, since the kinetics are

roughly logarithmic until the saturation point, and neither the wet height nor the friction

changes appreciably upon further UV irradiation, we conjecture that relatively few PFPA

units are bound to the dextran chains, and attempt to estimate the number of attachment

points from the measured values.

Assuming that the chain is attached to the PFPA layer at equidistant points along the chain,

and that these can be treated as single attachments (therefore neglecting trains), then the film

can be treated as a “brush” consisting of two end tails and a number of loops in the middle,

0

2

4

6

8

1 10 100

Laye

r p

erc

en

t d

extr

an

UV exposure, seconds

(C)

Page 75: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

74

with the spacing between points such that the tail and loop lengths are equal, as shown in the

Supporting Information. For example, if a 2 MDa chain is bound at two points, then that

chain can be treated as two 500 kDa tails and a 1MDa loop, which approximates to four

equally spaced 500 kDa end-attached chains. If these effective 500 kDa chains are swollen in

liquid to their bulk hydrodynamic radius,29

the theoretical film thickness can be calculated

from the measured dextran surface density, and compared to the measured value. Replacing

the single-point attachments with short trains would not significantly influence the result. The

best agreement with the measured wet thickness of the film was achieved when the 2 MDa

chain was bound at three points, yielding two 333 kDa tails and two 666 kDa loops, which if

equated to an effective brush of six 333 kDa chains would have a theoretical film thickness of

125.5 nm at the densest end of the gradient. While there are numerous sources of error in this

calculation, it supports the hypothesis that each chain is bound by only a few PFPA units. A

uniform 333 kDa brush of the measured density would have an L/2Rg value of 0.21, taking Rg

to be 19 nm.30

5.4.2 Friction measurements

The friction coefficient of the PAAm-g-PFPA-dextran films is highly load dependent, with

all but the sparsest films exhibiting a low-friction regime at low loads. A transition to a high-

friction regime occurs above a critical normal load, as shown in Figure 5.4. This behavior is

similar to that observed in PLL-g-dextran brushes,17,18

and Figures 5.5 and 5.6 show the

results from that study alongside the results of the PAAm-g-PFPA-dextran findings for

comparison. Figure 5.5 shows the friction coefficient in the low- and high-load regimes

plotted against the dextran monomer density, while Figure 5.6 shows the critical normal load

at which the friction increases, also plotted against the dextran monomer density. All

measurements were performed against a bare borosilicate sphere as well as a PLL-g-dextran-

coated sphere.

Page 76: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

75

Figure 5.4 Typical friction vs. load curves of a colloidal probe sliding on a PFPA-dex-coated

wafer exposed to UV for 5 seconds (yielding a coverage of 9 dextran monomers/nm2, white),

20 seconds (24 monomers/nm2, grey), and 120 seconds (35 monomers/nm

2, black). Fits of the

linear parts of each series, excluding the non-linear intermediate points, are extrapolated to

their intersecting point, the x-coordinate thereof representing the transition load.

0

50

100

150

200

0 50 100 150 200 250

Late

ral f

orc

e, n

N

Normal force, nN

Page 77: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

76

Figure 5.5 Coefficient of friction of PLL-g-dextran17

(black and white) and PAAm-g-PFPA-

dextran (blue) plotted against surface density of dextran monomers in both low and high load

0

1

2

3

4

5

6

0 10 20 30 40

Co

eff

icie

nt

of

fric

tio

n

Dextran monomers per nm2

(A)

0

0.1

0.2

0.3

0.4

0.5

0 10 20 30 40

Co

eff

icie

nt

of

fric

tio

n

Dextran monomers per nm2

(B)

Page 78: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

77

regimes (A) and just the low load regimes (B). The squares represent a bare borosilicate

sphere at low load (white or light blue) and high load (black or dark blue). The circles

represent a PLL-g-dextran coated borosilicate sphere at low load (white or light blue) and

high load (black or dark blue).

Figure 5.6 Critical load associated with the transition from low to high friction regimes of

PLL-g-dextran17

(grey) and PAAm-g-PFPA-dextran (blue) plotted against surface density of

dextran monomers. The squares represent a bare borosilicate sphere, while the circles

represent a PLL-g-dextran coated borosilicate sphere.

The PAAm-g-PFPA-dextran system, consisting of a more disordered loop-train-tail

architecture, showed a higher friction coefficient at low loads compared to the brush formed

by PLL-g-dextran, despite the greater wet film thickness and much higher surface monomer

density. However, the disordered system showed a lower friction coefficient at high loads

compared to the brush system, with the coefficient of friction decreasing as the dextran

density increased, while the opposite was true for the brush. The transition from low to high

friction regimes also occurred at much higher dextran surface monomer concentrations in the

disordered case than for the brush, indicating that the disordered system is less able to support

0

50

100

150

200

250

0 10 20 30 40

Tran

siti

on

Lo

ad, n

N

Dextran monomers per nm2

Page 79: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

78

high loads. Finally, the impact of dextran-coating the borosilicate sphere is opposite in the

high-friction regime, with the dextran coating exacerbating friction in the brush case while

reducing it in the disordered case.

The significance of the brush conformation has been demonstrated previously by collapsing

brushes with minimal changes to the solvent,31

which dramatically changes the mechanical

properties of the film. In the absence of significant hydrodynamic forces, longer, more

disordered systems will significantly interpenetrate and interact under even low loads. Brush

interpenetration increases friction even when inter-chain interactions from opposing

substrates, such as tangling or hydrogen bonding, are neglected.32

Tangling is also known to

increase friction coefficient between polymer brushes,33-36

although tangling is not possible

for the 5 kDa dextran chains, since they are too short and stiff. However, the interpenetration

of chains under high loads, with the subsequent formation of inter-chain hydrogen bonds,

nevertheless creates high friction between the surfaces.17

The impact of interpenetration,

entanglement and hydrogen bonding together, as in the case of the dextran loop-train-tail

system, is more complex. While hydrogen bonding between chains on the same substrate

may inhibit entanglement with chains of the counterface at low loads, the application of

higher loads and lateral forces does ultimately yield increased friction, which is controlled by

inter-chain association as well as tangling.37

The combination of increased chain entanglement and hydrogen bonding between the

counterfaces can explain the differences between the disordered, thicker dextran system and

the brush system. The friction coefficient of the brush is lower at low loads, due to its density

and the lack of tip interpenetration, thanks to the inter-chain hydrogen bonding holding the

chain ends in place. As the load increases, the ordered brush system is more capable of

resisting this interpenetration than the disordered system, which more rapidly becomes

enmeshed with the counterface. However, the ordering and hydrogen bonding show the

opposite effect once interpenetration is achieved, with interactions between opposing surfaces

of ordered chains creating much more resistance than is the case for the longer, disordered

conformation, where the chains are freer to rearrange themselves.

While these findings are significant in the field of brush and thin-film design, they are most

pertinent to the case of more complex architectures, in which gels and brushes may be

combined in various ways to achieve tailored properties. For example, it has been shown that

Page 80: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

79

short polymer brushes on top of gels reduce their friction, while long polymer brushes

increase it,38

a phenomenon that seems likely to be related to the differences outlined here.

5.5 Conclusions

The lubricating properties of a dextran surface-density gradient, attached at random intervals

to the surface by means of an intermediate azide-bearing monolayer, were investigated by

means of colloidal-probe lateral force microscopy. The results were contrasted with previous

work involving density gradients of end-attached dextran (a polymer brush density gradient).

Both systems showed highly load-dependent friction, with the randomly attached dextran

chains exhibiting less extreme differences than the polymer-brush case. For the randomly

attached system, significantly higher surface concentrations of dextran were necessary to

reach the same critical load (marking the transition from low to high friction regimes). This

suggests that the polymer-brush architecture significantly enhanced the load-bearing

properties of the dextran film. Increasing the density of the disordered dextran film reduced

the friction coefficient at high loads, while the opposite was true for the polymer brush. These

findings demonstrate that the brush architecture is critical in resisting collapse under applied

load, but is susceptible to inter-chain interactions between the surfaces, which increase

friction once the load has exceeded a critical threshold. This highlights the tradeoff involved

in designing film architectures to maximize layer thickness and lubricity under extreme loads

or to aim for an ideal brush with ultra-low friction at low and moderate loads. Future work on

heterogeneous architectures may allow tailored films that exhibit the appropriate mixture of

properties, depending on the expected loading of the film.

5.6 Supplementary information

Liquid VASE Fit Layers Thickness Refractive Index

Dextran Measured 1.51 + 0.01/λ2

PAAm-g-PFPA Measured before

spin-coating 1.45 + 0.01/λ

2

Silicon oxide Measured before

adsorption

Built in to

software

Silicon substrate Infinite Built in to

software

Page 81: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

80

Table 5.1 VASE multilayer model used to fit the dry measurements. As the refractive index

formula indicates, the PAAm-g-PFPA and dextran layers are treated as Cauchy films.

Liquid VASE Fit Layers Thickness Refractive Index

Ambient water Infinite 1.333

Dextran/water mixture (Bruggeman effective

medium) Measured

Measured mix of

1.51 and 1.333

PAAm-g-PFPA Measured dry

before spin-coating 1.45 + 0.01/λ

2

Silicon oxide Measured dry

before adsorption

Built in to

software

Silicon substrate Infinite Built in to

software

Table 5.2 VASE multilayer model used to fit the liquid-cell measurements. As the refractive

index formula indicates, the PAAm-g-PFPA layer is treated as a Cauchy film.

Page 82: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

81

Figure 5.7 Schematic representation of the model used to calculate probable dextran

attachment density. The 2 MDa dextran chain is attached to the substrate at either one, two,

or three points, spaced such that the end and loop lengths are equal. These are then treated as

discrete end-attached chains, and the height of the polymer brush of corresponding chain

length and surface density is then calculated assuming that the chains are swollen to their

hydrodynamic volume.

5.7 References

1. Wuge H. Briscoe, Simon Titmuss, Fredrik Tiberg, Robert K. Thomas, Duncan J.

McGillivray and Jacob Klein, “Boundary lubrication under water”, Nature, 2006,

10.1038/nature05196

Page 83: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

82

2. Duncan Dowson, “History of Tribology”, Professional Engineering Publishing, London,

1998.

3. Gregory L. Kenausis, Janos Vörös, Donald L. Elbert, Ningping Huang, Rolf Hofer,

Laurence Ruiz-Taylor, Marcus Textor, Jeffrey A. Hubbell and Nicholas D. Spencer,

“Poly(L-lysine)-g-Poly(ethylene glycol) Layers on Metal Oxide Surfaces: Attachment

Mechanism and Effects of Polymer Architecture on Resistance to Protein Adsorption”,

Journal of Physical Chemistry B, 2000, 10.1021/jp993359m

4. Ning-Ping Huang, Roger Michel, Janos Voros, Marcus Textor, Rolf Hofer, Antonella

Rossi, Donald L. Elbert, Jeffrey A. Hubbell and Nicholas D. Spencer, “Poly(L-lysine)-g-

poly(ethylene glycol) Layers on Metal Oxide Surfaces: Surface-Analytical

Characterization and Resistance to Serum and Fibrinogen Adsorption”, Langmuir, 2001,

10.1021/la000736+

5. Seunghwan Lee and Nicholas D. Spencer, Chapter 21 in Superlubricity, Elsevier Science

B.V., Amsterdam, 2007

6. Seunghwan Lee, Markus Müller, Monica Ratoi-Salagean, Janos Vörös, Stéphanie

Pasche, Susan M. De Paul, Hugh A. Spikes, Marcus Textor and Nicholas D. Spencer,

“Boundary Lubrication of Oxide Surfaces by Poly(L-lysine)-g-poly(ethylene glycol)

(PLL-g-PEG) in Aqueous Media”, Tribology Letters, 2003, 10.1023/A:1024861119372

7. Scott T. Milner, “Polymer Brushes”, Science, 1991, 10.1126/science.251.4996.905

8. Duncan Dowson, “Bio-tribology”, Faraday Discussions, 2012, 10.1039/C2FD20103H

9. Bruno Zappone, Marina Ruths, George W. Greene, Gregory D. Jay and Jacob N.

Israelachvili, “Adsorption, Lubrication, and Wear of Lubricin on Model Surfaces:

Polymer Brush-Like Behavior of a Glycoprotein”, Biophysical Journal, 2007,

10.1529/biophysj.106.088799

Page 84: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

83

10. Seunghwan Lee and Nicholas D. Spencer, “Sweet, Hairy, Soft, and Slippery”, Science,

2008, 10.1126/science.1153273

11. Surangkhana Martwiset, Anna E. Koh and Wei Chen, “Nonfouling Characteristics of

Dextran-Containing Surfaces”, Langmuir, 2006, 10.1021/la061064b

12. Wouter T. E. Bosker, Katja Patzsch, Martien A. Cohen Stuart and Willem Norde, “Sweet

brushes and dirty proteins”, Soft Matter, 2007, 10.1039/B618259C

13. Chiara Perrino, Seunghwan Lee, Sung Won Choi, Atsushi Maruyama and Nicholas D.

Spencer, “A Biomimetic Alternative to PEG as an Antifouling Coating: Resistance to

Non-Specific Protein Adsorption of Poly(L-lysine)-graft-Dextran”, Langmuir, 2008,

10.1021/la800947z

14. Chiara Perrino, Seunghwan Lee and Nicholas D. Spencer, “Surface-grafted Sugar Chains

as Aqueous Lubricant Additives: Synthesis and Macrotribological Tests of Poly(L-

lysine)-graft-Dextran (PLL-g-dex) Copolymers”, Tribology Letters, 2009,

10.1007/s11249-008-9402-6

15. Anwarul Ferdous, Hiromitsu Watanabe, Toshihiro Akaike and Atsushi Maruyama,

Poly(L-lysine)–graft–dextran copolymer: amazing effects on triplex stabilization under

physiological pH and ionic conditions (in vitro), Nucleic Acids Research, 1998, 26, 3949

16. Atsushi Maruyama, Hiromitsu Watanabe, Anwarul Ferdous, Maiko Katoh, Tsutomu

Ishihara and Toshihiro Akaike, “Characterization of interpolyelectrolyte complexes

between double-stranded DNA and polylysine comb-type copolymers having hydrophilic

side chains”, Bioconjugate Chemistry, 1998, 10.1021/bc9701510

17. Kenneth J. Rosenberg, Tolga Goren, Rowena Crockett and Nicholas D. Spencer, “Load-

induced transitions in the lubricity of adsorbed poly(L-lysine)-g-dextran as a function of

polysaccharide chain density”, ACS Applied Materials and Interfaces, 2011,

10.1021/am200521m

Page 85: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

84

18. Tolga Goren, Rowena Crockett and Nicholas D. Spencer, “Influence of Solutes on

Hydration and Lubricity of Dextran Brushes”, CHIMIA, 2012, 10.2533/chimia.2012.192

19. Stéphanie Pasche, Susan M. De Paul, Janos Vörös, Nicholas D. Spencer and Marcus

Textor, “Poly(L-lysine)-graft-poly(ethylene glycol) Assembled Monolayers on Niobium

Oxide Surfaces: A Quantitative Study of the Influence of Polymer Interfacial

Architecture on Resistance to Protein Adsorption by ToF-SIMS and in Situ OWLS”,

Langmuir, 2003, 10.1021/la034111y

20. Ângela Serrano, Olof Sterner, Sophie Mieszkin, Stefan Zürcher, Samuele Tosatti,

Maureen E. Callow, James A. Callow and Nicholas D. Spencer, “Nonfouling Response

of Hydrophilic Uncharged Polymers”, Advanced Functional Materials, 2013,

10.1002/adfm.201203470

21. Olof Sterner, Ângela Serrano, Sophie Mieszkin, Stefan Zürcher, Samuele Tosatti,

Maureen E. Callow, James A. Callow and Nicholas D. Spencer, “Photochemically

Prepared, Two-Component Polymer-Concentration Gradients”, Langmuir, 2013,

10.1021/la402168z

22. E. Palik, “Handbook of optical constants of solids”, New York Academic, Orlando,

1985.

23. D.A.G. Bruggeman, “Berechnung verschiedener physikalischer Konstanten von

heterogenen Substanzen”, Ann. Phys. (Leipzig), 1935, 24, 636-679

24. John E. Sader and Christopher P. Green, “In-plane deformation of cantilever plates with

applications to lateral force microscopy”, Review of Scientific Instruments, 2004,

10.1063/1.1667252

25. Ernst Meyer, “Atomic Force Microscopy”, Progress in Surface Science, 1992,

10.1016/0079-6816(92)90009-7

26. S. M. Cook, T. E. Schäffer, K. M. Chynoweth, M. Wigton, R. W. Simmonds and

Kristine M. Lang, “Practical implementation of dynamic methods for measuring atomic

Page 86: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

85

force microscope cantilever spring constants”, Nanotechnology, 2006, 10.1088/0957-

4484/17/9/010

27. John E. Sader, James W. M. Chon and Paul Mulvaney, “Calibration of rectangular

atomic force microscope cantilevers”, Review of Scientific Instruments, 1999,

10.1063/1.1150021

28. Rachel J. Cannara, Michael Eglin and Robert W. Carpick, “Lateral force calibration in

atomic force microscopy: A new lateral force calibration method and general guidelines

for optimization”, Review of Scientific Instruments, 2006, 10.1063/1.2198768

29. Kirsti A Granath, “Solution properties of branched dextrans”, Journal of Colloid Science,

1958, 10.1016/0095-8522(58)90041-2

30. Jonathan K. Armstrong, Rosalinda B. Wenby, Herbert J. Meiselman and Timothy C.

Fisher, “The Hydrodynamic Radii of Macromolecules and Their Effect on Red Blood

Cell Aggregation”, Biophysical Journal, 2004, 10.1529/biophysj.104.047746

31. Omar Azzaroni, Sergio Moya, Tamer Farhan, Andrew A. Brown and Wilhelm T. S.

Huck, “Switching the properties of polyelectrolyte brushes via “hydrophobic collapse””,

Macromolecules, 2005, 10.1021/ma051549r

32. Akihiro Nomura, Kenji Okayasu, Kohji Ohno, Takeshi Fukuda and Yoshinobu Tsujii,

“Lubrication Mechanism of Concentrated Polymer Brushes in Solvents: Effect of

Solvent Quality and Thereby Swelling State”, Macromolecules, 2011,

10.1021/ma200340d

33. Jacob Klein, in “Fundamentals of Tribology and Bridging the Gap Between the Macro-

and Micro/Nanoscales”, Kluwer Academic Pub., Dordrecht, 2001, p. 177.

34. Uri Raviv, Suzanne Giasson, Nir Kampf, Jean-François Gohy, Robert Jérôme and Jacob

Klein, “Lubrication by charged polymers”, Nature, 2003, 10.1038/nature01970

Page 87: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

86

35. Benoît Liberelle and Suzanne Giasson, “Friction and Normal Interaction Forces between

Irreversibly Attached Weakly Charged Polymer Brushes”, Langmuir, 2008,

10.1021/la702367f

36. Florent Goujon, Patrice Malfreyt and Dominic J. Tildesley, “Mesoscopic simulation of

entangled polymer brushes under shear: compression and rheological properties”,

Macromolecules, 2009, 10.1021/ma9000429

37. David M. Loveless, Nehal I. Abu-Lail, Marian Kaholek, Stefan Zauscher and Stephen L.

Craig, “Reversibly Cross-Linked Surface-Grafted Polymer Brushes”, Angewandte

Chemie, 2006, 10.1002/ange.200602508

38. Yutaka Ohsedo, Rikiya Takashina, Jian Ping Gong and Yoshihito Osada, “Surface

Friction of Hydrogels with Well-Defined Polyelectrolyte Brushes”, Langmuir, 2004,

10.1021/la036211+

Page 88: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

87

Chapter 6: Hydration and conformation in

dextran and PEG brushes

This work is from a manuscript in preparation. The ATR-IR experiments on PLL-g-dextran

and the related discussion were done by myself and Dr. Venkataraman Nagayanallur working

together, while the ATR-IR experiments on PLL-g-PEG were done solely by him. He also

contributed to the manuscript text and to discussions about the project. The rest of the

described work was done by myself. Dr. Rowena Crockett and Prof. Nicholas D. Spencer

supervised the work, contributed in discussions about the project steps, and corrected the

manuscript. Dr. Patrick Ilg also gave valuable suggestions concerning the simulation study.

6.1 Abstract

The hydrated conformation of dextran brushes at different chain densities was investigated

via optical-thickness measurements and course-grained molecular dynamics simulations. The

dextran conformation was controlled to a large degree by inter-chain interactions, with the

brush height being only weakly dependent on brush grafting density, in contrast to the scaling

law for ideal brushes in good solvent. To compare this behavior to the influence of hydration

on chain conformation, the variation of PEG and dextran brush structure under dry, humid

and wet conditions was studied via ATR-IR. Significant shifts in the C-O-C stretching bands

were observed in PEG brushes with changing humidity, indicative of conformational changes

due to the presence of water, while no corresponding shifts were observed in dextran. These

observations of the differing roles of water and structure in PEG and dextran brushes are

useful in interpreting their differing properties, as reported in the literature.

6.2 Introduction

Surfaces bearing densely packed polymers in good solvents are protected to an extraordinary

degree from adsorbates and foreign bodies. The steric barrier presented by the polymer chains

and their associated solvent molecules, as well as the osmotic pressure induced by the

favorability of interactions with the solvent, lend them characteristic low-friction and

nonfouling behavior, which can be observed over different chain chemistries and

architectures.1,2

Polymer brushes composed of biocompatible polymers such as poly(ethylene

Page 89: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

88

glycol) (PEG) are of interest as biocompatible and environmentally friendly non-fouling

surfaces, aqueous lubricants, and “stealth coatings” against immune system response.3-5

Polysaccharide brushes and thin films have also been produced,6-8

many incorporating

dextran – a naturally-occurring glucan consisting of β(16) glycosidic linkages, which

confer a relatively high degree of flexibility and inhibit the formation of stable crystalline

structures. PLL-g-dextran consists of dextran chains end-grafted to some of the pendant

lysine groups of poly(L-lysine).9-12

The remaining unreacted lysine groups are positively

charged under aqueous conditions, and thus the molecule adsorbs spontaneously onto

negatively charged surfaces, such as silicon dioxide, with sufficient chain density to force the

hydrophilic dextran chains into a brush conformation. PLL-g-dextran-based brushes behave

similarly to PLL-g-PEG in terms of adsorption behavior, fouling resistance, and lubricity

under low loads in aqueous conditions.11,12

However, the ability of the dextran chains to

hydrogen bond together influences the inter-chain interactions, as well as the hydration, of

the brush.

A widely cited model13

describes a helical conformation of PEG within a cage of water, with

the coincidence of the matching structures of PEG and the hydration cage lending high

solubility to the chemically amphiphilic polymer.14

Similar behavior has been observed in

PEG brushes, as discussed by Heuberger et al. and references therein.15

In contrast, glucose is

chemically akin to a water cluster, aside from the small hydrophobic patches on the two

carbohydrate ring faces, although the crystalline structure of certain glucans, such as cellulose

or amylose inhibits their solubility, and even dextran can form insoluble agglomerates in

water in some cases.16-18

In other words, PEG is chemically amphiphilic but structurally

water-soluble, while glucans are chemically hydrophilic but can be structurally water-

insoluble. PEG brushes also show decreasing hydration on a per-monomer basis with

increasing chain density,19

whereas dextran brushes exhibit no correlation between hydration

and chain density.20

Infrared spectroscopy is a sensitive tool to determine molecular structure and chain

conformation of polymers. Several structural investigations of polysaccharides, including

dextran, in their bulk phase as well as the changes occurring in solutions have been probed

using IR spectroscopy.21,22

However, structural investigations of polysaccharides grafted to

surfaces and the conformational changes occurring upon hydration/dehydration have, to the

best of our knowledge, not been reported.

Page 90: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

89

Computer simulations can aid in the understanding of dextran chain conformation and

solvation, but atomistic simulations of brushes with explicit water are too computationally

intensive to study over useful timescales. Therefore, a coarse-grained force field was used to

represent the dextran and water molecules. The M3B model was first developed for 1→4-

linked oligosaccharides in water,23

and subsequently modified to represent 1→6-linked

dextran chains.24

The model treats each dextran monomer as three beads corresponding to

carbons 1, 4, and 6, with water molecules represented by a single bead. Intermolecular

bonding is limited to two-body Morse functions, although the model parameterization was

performed using atomistic simulations that did include hydrogen bonding.23

6.3 Materials and methods

6.3.1 Poly(L-lysine)-graft-dextran (PLL-g-dextran) and poly(L-

lysine)-graft-poly(ethylene glycol) (PLL-g-PEG)

The poly(L-lysine)-graft-dextran (PLL-g-dextran) copolymer used in these experiments was

synthesized during previous work.11,12

Briefly, dextran (dextran T5, 5.2 kDa, polydispersity

1.8, Pharma-cosmos A/S, Denmark) undergoes reductive amination with poly(L-lysine)-HBr

(20 kDa, polydispersity 1.1, Sigma-Aldrich, Switzerland) in a sodium borate buffer (0.1 M,

pH 8.5, 0.4 M NaCl). The terminal dextran aldehyde group reacts with the primary amine

groups of PLL to form a Schiff base, which is subsequently reduced by sodium

cyanoborohydride (NaBH3CN). The resultant PLL-g-dextran graft copolymer is separated

from the unreacted materials via ultracentrifugation. The grafting ratio of lysine monomers to

dextran chains can be controlled during synthesis by adjusting the reactant ratio, and was

subsequently determined through 1H-NMR spectroscopy of PLL-g-dextran in D2O and

elemental analysis. The particular copolymer studied here was PLL(20)-g[5.3]-dextran(5),

where 20 kDa is the molar mass of PLL including the Br- counterions, 5.3 is the grafting

ratio, and 5 kDa is the weight-average molecular weight (Mw) of the dextran. On average,

these dextran chains consist of 32 monomer units, which, assuming each monomer to be 0.7

nm long, corresponds to a fully stretched chain length of 22.3 nm.

The poly(L-lysine)-graft-poly(ethylene glycol) (PLL-g-PEG) copolymer used in these

experiments was obtained from SuSoS (SuSoS AG, Dübendorf, Switzerland). The particular

copolymer studied here was PLL(20)-g(3.5)-PEG(2), where 20 kDa is the molar mass of PLL

Page 91: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

90

including the Br- counterions, 3.5 is the grafting ratio, and 2 kDa is the weight-average

molecular weight (Mw) of the PEG chains. On average, these PEG chains consist of 45

monomer units, which, assuming each monomer to be 0.36 nm long, corresponds to a fully

stretched chain length of 16.3 nm.

6.3.2 Transmission interference adsorption sensing (TInAS)

Transmission interference adsorption sensing (TInAS) is an optical biosensing technique that

monitors interference fringes formed when light is transmitted through a film adsorbed onto a

special multilayered TInAS sensor.25

The wavelengths of the interference peaks are

correlated to the optical thickness of the adsorbed layer, while the peak shapes may be used

to measure the refractive index.26

In a typical TInAS experiment, an SiO2-coated sensor is

exposed on one side to a solution containing the desired adsorbate, and the optical thickness

of the deposited film is measured in real time, with a sensitivity on the order of 1 ng/cm2

(when converted to adsorbed density) and a temporal resolution below 1 s.25

All spectra were

fitted using both the standard TInAS software25

and the TInAS-FIT software package.26

From

these data, the film’s optical mass may be calculated using the de Feijter method.27

TInAS sensors were produced by sputter coating flat, clean glass substrates with 25 nm Al,

followed by 2-3 μm SiO2, resulting in a surface with ~8 nm RMS roughness. Prior to use,

these sensors were cleaned by sonicating twice in toluene, then twice in 2-propanol, each for

five minutes. Afterwards, the sensors were exposed to O2 plasma for two minutes (Plasma

Cleaner/Sterilizer, PDC-32G instrument, Harrick, Ossining, NY, USA). After mounting the

sensors in the flow cell, the baseline and adsorption measurements were performed in

ultrapure water. A polymer concentration of 0.02 g/L was used for adsorption, with a

continuous flow rate of 20 μL/min to replenish the solution without affecting adsorption

kinetics.

6.3.3 Coarse-grained dextran simulations

The goal of the simulation was to determine the equilibrium conformation of the hydrated

dextran brush at varying adsorbed densities. To this end, a coarse-grained force field was

used to represent the dextran and water molecules.23,24

Unfortunately, due to the

conformational constraints imposed by the triangular geometry of the coarse-grained dextran

chains, equilibrated polymer conformations are difficult to achieve. Starting from straight

Page 92: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

91

conformations, molecular dynamics simulations were performed, following an annealing

procedure to equilibrate the system.

Molecular dynamics simulations in this study were performed in the NVT ensemble, with the

equations of motion being integrated using the Verlet algorithm and a time step of 1.0 fs. All

simulations were performed with LAMMPS.28

The system temperature was set at 300 K and

controlled with the Nose-Hoover thermostat. The simulation floor and ceiling were bounded

by Lennard-Jones walls, while the lateral dimensions were periodic. The simulation consisted

of six chains, each 32 monomers long, with the anchoring monomers fixed in place 3.5 Å

above the minimum of the simulation floor wall. The simulation box height was set to 100 Å

for the sparser chains and 120 Å for the densest chains – values that were determined using

trial runs to be sufficient to prevent the ceiling from playing a role in the brush equilibration.

The lateral dimensions of the box were set by the desired chain density of the simulation. The

dextran monomers were placed with random x and y coordinates, with the z-coordinate

increasing regularly from the anchoring end at 3.5 Å up to a maximum value of 35 Å, 65 Å,

or 95 Å, in order to sample different starting heights. Water molecules were randomly placed

throughout the cell to achieve a density of one particle per 33 cubic Ångstroms. Energy

minimization using the Polak-Ribiere conjugate gradient algorithm was first performed on

these randomized positions to achieve a random, but more physically reasonable, initial

condition, following which the annealing procedure was applied. The system was heated

from 300 K to 500 K over 0.1 ns, then cooled back down to 300 K over another 0.1 ns, after

which another energy minimization was performed. This procedure was repeated for 2-4 ns.

After each energy minimization, the system energy and the average chain height were

monitored to ensure that a steady state was reached within the first half of the total simulation

time.

To validate this protocol, a separate series of simulations was performed to measure the Rg of

a single dextran chain in explicit water. A single, unattached chain of 5 kDa or 35 kDa was

simulated in a fully periodic box according to the same procedure as described above, at the

end of which the Rg of the molecule was calculated. The results for both chain lengths were

in excellent agreement with experimental measurements of dextran Rg.29

6.3.4 ATR crystal preparation and brush attachment

ATR-IR spectra were measured on a Nicolet 5700 FTIR spectrometer equipped with a liquid-

nitrogen-cooled MCT-A detector and multiple-internal-reflection accessory (Perkin-Elmer)

Page 93: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

92

mounted with a liquid cell. 52x20x2mm 45° trapezoidal germanium ATR prisms (Crystran,

Dorset, UK) were employed for the measurement of the ATR spectra of PLL-g-dextran and

PLL-g-PEG adsorbed on the native oxide surface of the Ge crystal. Spectra of the adsorbed

polymer were measured both in a dry and a wet state by referencing to a freshly plasma-

cleaned Ge ATR crystal and a clean crystal in contact with ultrapure (MilliQ) water,

respectively. Measurements at different relative humidities were performed using a custom-

built, two-way flow system.

For the measurements of PLL-g-dextran, the crystal was cleaned by sonication in COBAS

cleaner (Roche), followed by ultrapure water. Next, both sides of the crystal were exposed to

O2 plasma for one minute (Plasma Cleaner/Sterilizer, PDC-32G instrument, Harrick,

Ossining, NY, USA). The thickness of the surface oxide layer was measured by ellipsometry

(described below), following which the crystal was exposed to a 0.1 M solution of PLL-g-

dextran for 30 minutes, and then rinsed with ultrapure water and dried under N2. The final

PLL-g-dextran thickness on the surface was then measured by ellipsometry.

Both the germanium oxide layer and PLL-g-dextran film thicknesses were measured in

ambient air by ellipsometry using a M-2000 V spectroscopic ellipsometer (J.A. Woollam Co.,

Inc., Lincoln, NE, USA) in reflection at 70° between 370-800 nm. The variable-angle

spectroscopic ellipsometry (VASE) spectra were analyzed using WVASE32 software. The

thickness of PLL-g-dextran was obtained by fitting to a multilayer model, consisting of a

bottom layer of germanium (using the built-in optical parameters), a layer of germanium

oxide with refractive index of 1.65 and the thickness measured prior to adsorption, and a

generalized Cauchy layer (A = 1.45, B = 0.01, C = 0).30

6.4 Results and discussion

6.4.1 TInAS thickness and refractive index measurements of

PLL-g-dex

TInAS sensors were exposed to an 0.02 M solution of PLL-g-dextran(5 kDa) in ultrapure

water, and the spectra were measured in real time as the brush was formed. The spectra were

then interpreted with the standard TInAS software,25

which measures the thickness of the top

film assuming a refractive index of 1.45, as well as the TInAS-FIT software package, which

measures both the thickness and the refractive index of the film using a different algorithm.26

Page 94: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

93

The two methods showed excellent agreement, reporting a maximum adsorbed mass of 306 ±

15 ng/cm2. The TInAS-FIT thickness and refractive index values have a higher mutual

uncertainty than their combined optical mass, and this uncertainty is reduced over the

measurement with increasing film thickness.26

The measurements indicated that after adsorption had slowed down to within the drift of the

instrument, the hydrated brush had a mean thickness of 5.6 + 0.4 nm and a refractive index of

1.41 + 0.2. A representative measurement is shown in Figure 6.1, with the calculated optical

mass of the same experiment shown in the Supporting Information. The final thickness of the

film was achieved within the first minute of adsorption, with the refractive index increasing

much more slowly, achieving its ultimate value only after approximately ten minutes.

Figure 6.1 Simultaneous thickness (blue) and refractive index (red) measurements of PLL-g-

dextran adsorption kinetics in the TInAS. The final film thickness is rapidly achieved, while

the refractive index grows more gradually.

The TInAS measurements showed that the PLL-g-dextran brush reaches its maximum

thickness very early in the adsorption curve, when the adsorbed chain density is still low.

This implies that the chain height is not significantly influenced by the packing, and the

dextran chains maintain their shape in solution instead of adopting a mushroom-like

conformation in the absence of steric forces from neighboring chains to force them into an

extended conformation. The ultimate adsorbed mass of 306 ng/cm2 corresponds to an average

1.33

1.35

1.37

1.39

1.41

1.43

1.45

1.47

1.49

1.51

0

1

2

3

4

5

6

7

8

9

10

0 5 10 15 20R

efr

acti

ve In

dex

Thic

kne

ss, n

m

Time, min

Page 95: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

94

inter-chain distance L of 1.9 nm, or a brush parameter L/2Rg of 0.5, taking the radius of

gyration Rg to also be 1.9 nm.31

Comparing this value to the simulation results below (Figure

6.2) shows a relatively sharp increase in chain stretching if this grafting density is exceeded,

which would likely also correspond to significant steric hindrance to further polymer

adsorption.

6.4.2 Simulations

Coarse-grained simulations of dextran brushes in explicit water were used to obtain brush

heights at varying chain densities and with different initial simulation heights. Figure 6.2

shows these heights plotted against the chain grafting density and against L/2Rg, the inter-

chain distance divided by the radius of gyration of 5 kDa dextran, taken to be 1.9 nm.31

Initial

brush heights of 3 nm, 6 nm, and 9 nm are shown for each brush density. The brush height

was defined as the z-coordinate bounding 90% of the dextran monomers. A visual

representation of this is shown in the Supporting Information.

The dextran brush height is seen to increase with increasing chain density. The simulation

initial conditions also have a statistically significant, but comparatively minor, influence on

the brush height. Averaging the different initial conditions at each chain density yields a

curve which is best fit by a power law with the scaling exponent of 0.234, much lower than

the expected value of 1/3 for a good solvent (significantly different with >99% confidence, as

detailed in the Supporting Information). The simulations show an average thickness of ~5.8

nm at the maximum experimental brush density of L/2Rg = 0.5, in reasonably good

agreement with the thickness measured in TInAS.

Page 96: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

95

Figure 6.2 (A) Equilibrium dextran brush height vs. chain density, obtained from coarse-

grained simulations using the modified M3B model. Starting brush heights of 3 nm (blue

diamonds), 6 nm (red squares), and 9 nm (green triangles) were used. The brush height was

defined as the z-coordinate bounding 90% of the dextran monomers. A power law best fit is

shown (solid line) corresponding to a scaling exponent of 0.234, with a 95% confidence

0

2

4

6

8

10

0 0.2 0.4 0.6 0.8 1 1.2

Equ

ilib

riu

m B

rush

He

igh

t, n

m

Grafting Density, nm-2

(A)

0

2

4

6

8

10

0 0.2 0.4 0.6 0.8 1 1.2

Equ

ilib

riu

m B

rush

He

igh

t, n

m

L/2Rg

(B)

Page 97: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

96

interval (dotted lines), corresponding to an exponent range of 0.175-0.283. (B) The same data

plotted against L/2Rg.

Density profiles of the dextran monomers as well as the surrounding water were obtained

from the simulations. A representative plot is shown in Figure 6.3. In every case, the water

concentration increased and the dextran density decreased with distance from the surface,

aside from binning artifacts. The binning interval used for all plots was 0.4 nm. The dextran

density profiles of each brush density, averaged over all initial conditions, are shown in

Figure 6.4. The density profiles of the sparser brushes are typical for moderately stretched

brushes,32-34

while the densest brushes exhibit more slab-like conformations.

Figure 6.3 Representative density profiles of a dextran brush (solid line) and the surrounding

water (dashed line), showing an increase in water concentration with decreasing dextran

concentration.

0

2

4

6

8

10

12

14

16

18

20

0

0.5

1

1.5

2

0 3 6 9 12W

ate

rs p

er

nm

3

Dex

tran

mo

no

me

rs p

er

nm

3

Distance from surface, nm

Page 98: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

97

Figure 6.4: Equilibrium dextran brush density profile for inter-chain distances of 1 nm

(blue), 1.5 nm (red), 2 nm (green), 3 nm (purple), and 4 nm (orange).

The simulation results also show a relatively weak relationship between the inter-chain

spacing and the equilibrium brush height. Theoretical and experimental studies show that

brushes in good solvents tend to grow proportionally to the cube root of the chain density in a

good solvent, and to higher powers with lesser solvent quality or with higher brush densities,

due to the increased interaction between monomers.35-37

In contrast, the best fit to the

simulation results has an exponent of only 0.24, and a fit with an exponent of 0.3 or higher is

excluded at the 99% confidence level.

One possible cause of a weak dependence of dextran brush height on chain density is chain

stiffness. SAXS measurements showed that dextran has a persistence length of over 2.75

nm.38

The ATR-IR spectroscopic data also shows that the PEG chains of the PLL-g-PEG

adsorbed on the surface are more sensitive to the presence of water molecules in the

surrounding medium (both in vapor phase or in liquid) than the corresponding dextran chains

in PLL-g-dextran adsorbed on the surface, as further described below. It should be noted that

studies of the scaling behavior of PEG brushes did not mention any discrepancies from

theory.39

0

0.5

1

1.5

2

2.5

0 2 4 6 8 10

Dex

tran

mo

no

me

rs p

er

nm

3

Distance from surface, nm

1

1.5

2

3

4

Page 99: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

98

Another possible factor contributing to the relatively constant conformation of dextran

brushes is inter-chain hydrogen bonding. The combination of inter-chain hydrogen bonding

and chain stiffness can cause the chains to cluster and extend further than the steric hindrance

of crowding alone would cause. Dextran brushes have been shown to resist collapse and

spreading under applied load and shear, while interpenetration of opposing dextran brushes

under shear leads to extremely high friction due to hydrogen bonding between the opposing

chains.40

The compressibility and shear moduli of 5 kDa dextran brushes in aqueous solutions

are significantly higher than those of 5 kDa PEG brushes as measured in SFA41

and QCM,42

respectively, despite the higher solubility of dextran than PEG in water. Hydrogen bonding

also causes short-chain dextran to form insoluble agglomerates in solution, which are

resistant even to aggressive sonication and heating.16-18

6.4.3 ATR-IR

VASE measurements of PLL-g-dextran adsorbed onto the Ge ATR crystal yielded an optical

layer thickness of 0.74 + 0.23 nm. The lower thickness compared to PLL-g-PEG(2) in other

studies (e.g 2 + 0.28 nm on Nb2O5)43

can be due to the lower charge density or the lower

stability of germanium oxide, compared to silica or niobia, which can impact adsorption of

the PLL moieties, as well as possibly changing in optical thickness during the adsorption

process compared to the prior baseline measurement. This variability was acceptable given

that the aim of this work was to qualitatively understand the differences between PEG and

dextran behavior, rather than to quantitatively compare PEG or dextran spectra under

different conditions.

Figures 6.5 and 6.6 display representative ATR-IR spectra of PLL-g-dextran and PLL-g-PEG

adsorbed on the native oxide layer of a Ge ATR crystal, respectively. Characteristic bands of

the dextran moieties are clearly distinguishable in the spectra in Figure 6.5. A comparison of

the powder spectrum of dextran and that of PLL-g-dextran adsorbed on Ge over the entire

spectral range (4000-900 cm-1

) is given in the supporting information.

The spectra in Figures 6.5 and 6.6 exhibit two sets of peaks that are readily assignable to CH2

bending modes and C-O, C-C stretching modes. The peaks due to the CH2 bending modes are

not well resolved in the spectra of dextran, in comparison to PEG. However, the C-O-C

stretching region of dextran exhibits several well-resolved peaks in contrast to that of PEG.

This qualitative difference in spectra between PEG and dextran is understandable considering

Page 100: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

99

the several distinct C-O, C-C linkages of the sugar units compared to PEG with only one type

of C-C and C-O bonds.

Figure 6.5 ATR-IR spectra of PLL-g-dextran adsorbed onto a native oxide layer of Ge ATR

crystal measured dry, humid and immersed in water. C-O-C stretching bands (combination of

C-O and C-C stretching modes) characteristic of different positions within the dextran

molecule are clearly distinguishable.

Page 101: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

100

Figure 6.6 ATR-IR spectra of PLL-g-PEG adsorbed onto a native oxide layer of Ge ATR

crystal measured dry, humid and immersed in water. CH2 bending modes and C-O-C

stretching bands (combination of C-O and C-C stretching modes) characteristic of the PEG

chains are clearly distinguishable.

Apart from the qualitative differences in the spectra of dextran and PEG, a comparison of the

spectra of the adsorbed polymers measured in dry, humid and immersed in water reveal some

further details. The spectra of PLL-g-dextran measured under dry (red) and humid (blue)

conditions are essentially identical. No change in either the positions or the intensities of the

C-O-C stretching bands are observable. The spectrum of adsorbed PLL-g-dextran measured

in water also shows a few minor changes. The C-O-C stretching band at around 1018 cm-1

in

the dry/humid condition is shifted to 1014 cm-1

in the spectrum measured in water. The

intensity of this band, relative to the band at 1040 cm-1

, increases in the wet condition in

comparison to the dry/humid state. These changes in the spectral intensity and positions are

most likely due to the conformational changes happening to the dextran chains upon

Page 102: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

101

hydration. However, the spectra of PLL-g-PEG measured under similar conditions exhibit

more significant changes in the positions of the conformationally sensitive C-O-C stretching

bands.

In the case of PLL-g-PEG (Figure 6.6), appreciable changes in the position of the C-O-C

stretching bands are already visible between the dry and humid conditions. The C-O-C

stretching band at 1103 cm-1

in the dry state (red) shifts to 1095 cm-1

(blue) under humid

(RH90%) conditions and further to 1087 cm-1

(black) in the spectrum measured in water.

These results are in general agreement with the previously published results on oligo- or

poly- ethylene glycol self-assembled monolayers (SAM). For example, Skoda et al. have

observed shifts of the order of 17 cm-1

in the position of C-O-C stretching bands for oligo-

(ethylene glycol) terminated alkanethiol SAMs on gold, measured in air and in water by

means of polarization-modulation IR spectroscopy.44

These changes in peak positions are

related to the penetration of water molecules into the adsorbed polymer and the hydrogen

bonding of the water molecules to the ether linkages, causing the vibrational frequencies to

shift to lower wavenumbers.

6.5 Conclusions

The hydrated conformation of dextran and PEG brushes were investigated by means of

optical-thickness measurements, computer simulations, and ATR-IR measurements. The IR

results showed that PEG conformation is influenced by the presence of water, with the C-O-C

stretching bands shifting in position between dry, humid, and wet conditions, unlike dextran,

which exhibits no major band shifts. The optical measurements and computer simulations

suggested that the dextran-brush conformation is instead strongly influenced by inter-chain

interactions. Optical measurements showed that brush height was largely independent of

grafting density over the experimental range, while the simulation results, which reached

higher grafting densities, showed a smaller increase in height than predicted by the scaling

law for ideal brushes in good solvent. These findings are in agreement with literature

regarding the presence of an equilibrium hydrated structure for PEG brushes, which if

partially dried induces measurable conformation changes in the PEG chains, as well as the

significant inter-chain hydrogen bonding in dextran brushes, which is responsible for their

characteristic resistance to collapse under applied load and shear.

Page 103: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

102

6.6 Supporting information

Figure 6.7 Optical mass calculated from the simultaneous thickness and refractive index

measurements shown in Figure 6.1, using the de Feijter equation shown below. For dextran in

water, the refractive index increment dn/dc is taken to be 0.15 cm3/g.

0

50

100

150

200

250

300

0 5 10 15 20

Op

tica

l Mas

s, n

g/c

m2

Time, min

0

0.5

1

1.5

2

2.5

0 2 4 6 8 10

Dex

tran

mo

no

me

rs p

er

nm

3

Distance from surface, nm

Page 104: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

103

Figure 6.8 Equilibrium dextran brush density profile for inter-chain spacings of 1 nm (blue),

1.5 nm (red), 2 nm (green), 3 nm (purple), and 4 nm (orange). The dotted lines indicate the z-

coordinate bounding the lower 90% of dextran monomers, which is taken as the brush height

for the purpose of this study.

Simulation statistics

Statistical tests were performed to ascertain whether the exponent of a power-law fit differed

significantly from the hypothetical value of 1/3, separately for each set of initial conditions

(starting heights of 3, 6, or 9 nm), taking into account the standard error of each measurement

as shown in Figure 6.4. The results are shown in Table 6.1.

Initial Height, nm Scaling Exponent Best Fit 99% Confidence Interval R2 P value

3 0.2416 0.1850 to 0.2982 0.8665 0.0002

6 0.2463 0.1997 to 0.2928 0.9093 < 0.0001

9 0.2144 0.1611 to 0.2677 0.863 < 0.0001

Table 6.1 Statistical tests of the scaling exponent of the data shown in Figure 6.2.

Figure 6.9 Representative IR spectra comparing the spectra of pure dextran (black), PLL-g-

dextran (red) measured as KBr pellets along with that of the PLL-g-dextran adsorbed onto a

Page 105: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

104

Ge ATR crystal surface (blue). The spectra in the fingerprint region, sensitive to the structural

changes of the dextran chains (1350-950 cm-1

), are considered in detail in Figure 6.5.

6.7 References

1. Seunghwan Lee and Nicholas D. Spencer, Chapter 21 in Superlubricity, Elsevier Science

B.V., Amsterdam, 2007

2. Uri Raviv, Suzanne Giasson, Nir Kampf, Jean-François Gohy, Robert Jérôme and Jacob

Klein, “Lubrication by charged polymers”, Nature, 2003, 10.1038/nature01970

3. Gregory L. Kenausis, Janos Vörös, Donald L. Elbert, Ningping Huang, Rolf Hofer,

Laurence Ruiz-Taylor, Marcus Textor, Jeffrey A. Hubbell and Nicholas D. Spencer,

“Poly(L-lysine)-g-Poly(ethylene glycol) Layers on Metal Oxide Surfaces: Attachment

Mechanism and Effects of Polymer Architecture on Resistance to Protein Adsorption”,

Journal of Physical Chemistry B, 2000, 10.1021/jp993359m

4. Ning-Ping Huang, Roger Michel, Janos Voros, Marcus Textor, Rolf Hofer, Antonella

Rossi, Donald L. Elbert, Jeffrey A. Hubbell and Nicholas D. Spencer, “Poly(L-lysine)-g-

poly(ethylene glycol) Layers on Metal Oxide Surfaces: Surface-Analytical

Characterization and Resistance to Serum and Fibrinogen Adsorption”, Langmuir, 2001,

10.1021/la000736+

5. Seunghwan Lee, Markus Müller, Monica Ratoi-Salagean, Janos Vörös, Stéphanie

Pasche, Susan M. De Paul, Hugh A. Spikes, Marcus Textor and Nicholas D. Spencer,

“Boundary Lubrication of Oxide Surfaces by Poly(L-lysine)-g-poly(ethylene glycol)

(PLL-g-PEG) in Aqueous Media”, Tribology Letters, 2003, 10.1023/A:1024861119372

6. Surangkhana Martwiset, Anna E. Koh and Wei Chen, “Nonfouling Characteristics of

Dextran-Containing Surfaces”, Langmuir, 2006, 10.1021/la061064b

Page 106: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

105

7. Wouter T. E. Bosker, Katja Patzsch, Martien A. Cohen Stuart and Willem Norde, “Sweet

brushes and dirty proteins”, Soft Matter, 2007, 10.1039/B618259C

8. Nolan B. Holland, Yongxing Qiu, Mark Ruegsegger and Roger E. Marchant,

“Biomimetic engineering of non-adhesive glycocalyx-like surfaces using oligosaccharide

surfactant polymers”, Nature, 1998, 10.1038/33894

9. Anwarul Ferdous, Hiromitsu Watanabe, Toshihiro Akaike and Atsushi Maruyama,

Poly(L-lysine)–graft–dextran copolymer: amazing effects on triplex stabilization under

physiological pH and ionic conditions (in vitro), Nucleic Acids Research, 1998, 26, 3949

10. Atsushi Maruyama, Hiromitsu Watanabe, Anwarul Ferdous, Maiko Katoh, Tsutomu

Ishihara and Toshihiro Akaike, “Characterization of interpolyelectrolyte complexes

between double-stranded DNA and polylysine comb-type copolymers having hydrophilic

side chains”, Bioconjugate Chemistry, 1998, 10.1021/bc9701510

11. Chiara Perrino, Seunghwan Lee, Sung Won Choi, Atsushi Maruyama and Nicholas D.

Spencer, “A Biomimetic Alternative to PEG as an Antifouling Coating: Resistance to

Non-Specific Protein Adsorption of Poly(L-lysine)-graft-Dextran”, Langmuir, 2008,

10.1021/la800947z

12. Chiara Perrino, Seunghwan Lee and Nicholas D. Spencer, “Surface-grafted Sugar Chains

as Aqueous Lubricant Additives: Synthesis and Macrotribological Tests of Poly(L-

lysine)-graft-Dextran (PLL-g-dex) Copolymers”, Tribology Letters, 2009,

10.1007/s11249-008-9402-6

13. Roland Kjellander and Ebba Florin, “Water structure and changes in thermal stability of

the system poly(ethylene oxide)–water”, Journal of the Chemical Society, Faraday

Transactions 1: Physical Chemistry in Condensed Phases, 1981, 10.1039/F19817702053

14. Mikael Bjoerling, Gunnar Karlstroem and Per Linse, “Conformational adaption of

poly(ethylene oxide): A carbon-13 NMR study”, Journal of Physical Chemistry, 1991,

10.1021/j100170a060

Page 107: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

106

15. Manfred Heuberger, Tanja Drobek and Nicholas D. Spencer, “Interaction Forces and

Morphology of a Protein-Resistant Poly(ethylene glycol) Layer”, Biophysical Journal,

2005, 10.1529/biophysj.104.045443

16. Donald E. Cadwallader Jr., Charles H. Becker, Joan H. Winters and David Marcus, “A

note on particulate matter encountered in some dextran injections”, Journal of the

American Pharmaceutical Association, 1958, 10.1002/jps.3030471219

17. Allene Jeanes, N. C. Schieltz and Carl Wilham, “Molecular association in dextran and in

blanched amylaceous carbohydrates”, Journal of Biological Chemistry, 1948, 176, 617-

627

18. Masuo Aizawa, Shuichi Suzuki, Tatuo Kuoka, Noriyasu Nakajima and Yutaka Iwao,

“Stability of Dextran Solutions”, Bulletin of the Chemical Society of Japan , 1976,

10.1246/bcsj.49.2061

19. Markus Müller, “Aqueous Lubrication by Means of Surface-bound Brush-like

Copolymers”, PhD Thesis No. 16030, ETH Zürich, 2005, 10.3929/ethz-a-005025141,

page 118

20. Chiara Perrino, “Poly(L-lysine)-g-dextran (PLL-g-dex)”, PhD Thesis No. 18224, ETH

Zürich, 2009, 10.3929/ethz-a-005804590, page 88

21. Anton N. J. Heyn, “The infrared absorption spectrum of dextran and its bound water”,

Biopolymers, 1974, 10.1002/bip.1974.360130304

22. Kirill I Shingel, “Determination of structural peculiarities of dexran, pullulan and γ-

irradiated pullulan by Fourier-transform IR spectroscopy”, Carbohydrate Research,

2002, 10.1016/S0008-6215(02)00209-4

23. Valeria Molinero and William A. Goddard III, “M3B: A Coarse Grain Force Field for

Molecular Simulations of Malto-Oligosaccharides and Their Water Mixtures”, Journal of

Physical Chemistry B, 2004, 10.1021/jp0354752

Page 108: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

107

24. X. Zhang, J.-C. Wang, K. M. Lacki and A. I. Liapis, “Construction by Molecular

Dynamics Modeling and Simulations of the Porous Structures Formed by Dextran

Polymer Chains Attached on the Surface of the Pores of a Base Matrix: Characterization

of Porous Structures”, Journal of Physical Chemistry B, 2005, 10.1021/jp053421h

25. Manfred Heuberger and Tobias E. Balmer, “The transmission interferometric adsorption

sensor”, Journal of Physics D: Applied Physics, 2007, 10.1088/0022-3727/40/23/S07

26. Takumi Sannomiya, Tobias E. Balmer, Manfred Heuberger and Janos Vörös,

“Simultaneous refractive index and thickness measurement with the transmission

interferometric adsorption sensor”, Journal of Physics D: Applied Physics, 2010,

10.1088/0022-3727/43/40/405302

27. J. A. De Feijter, J. Benjamins and F. A. Veer, “Ellipsometry as a tool to study the

adsorption behavior of synthetic and biopolymers at the air–water interface”,

Biopolymers, 1978, 10.1002/bip.1978.360170711

28. Steve Plimpton, “Fast Parallel Algorithms for Short-Range Molecular Dynamics”,

Journal of Computational Physics, 1995, 10.1006/jcph.1995.1039

29. Sabine M. Görisch, Malte Wachsmuth, Katalin Fejes Tóth, Peter Lichter and Karsten

Rippe, “Histone acetylation increases chromatin accessibility”, Journal of Cell Science,

2005, 10.1242/jcs.02689

30. E. Palik, “Handbook of optical constants of solids”, New York Academic, Orlando,

1985.

31. Malvern Zetasizer Nano application note MRK839-01, 2007

32. Gary S. Grest, “Normal and Shear Forces Between Polymer Brushes”, Advances in

Polymer Science, 1999, 10.1007/3-540-69711-X_4

33. Fang Yin, Dmitry Bedrov, Grant D. Smith and S. Michael Kilbey II, “A Langevin

dynamics simulation study of the tribology of polymer loop brushes”, Journal of

Page 109: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

108

Chemical Physics, 2007, 10.1063/1.2757620

34. Patrick S. Doyle, Eric S. G. Shaqfeh and Alice P. Gast, “Rheology of Polymer Brushes:

A Brownian Dynamics Study”, Macromolecules, 10.1103/PhysRevLett.78.1182

35. E. B. Zhulina, O. V. Borisov, V. A. Pryamitsyn and T. M. Birshtein, “Coil-globule type

transitions in polymers. 1. Collapse of layers of grafted polymer chains”,

Macromolecules, 1991, 10.1021/ma00001a023

36. Avraham Halperin, “Collapse of grafted chains in poor solvents”, Journal de Physique,

1988, 10.1051/jphys:01988004903054700

37. Lionel C. H. Moh, Mark D. Losego and Paul V. Braun, “Solvent Quality Effects on

Scaling Behavior of Poly(methyl methacrylate) Brushes in the Moderate- and High-

Density Regimes”, Langmuir, 2011, 10.1021/la2002139

38. Sunil K. Garg and Salvatore S. Stivala, “Assessment of Branching in Polymers from

Small-Angle X-Ray Scattering (SAXS)”, Journal of Polymer Science: Polymer Physics

Edition, 1978, 10.1002/pol.1978.180160808

39. Stéphanie Louguet, Anitha C. Kumar, Nicolas Guidolin, Gilles Sigaud, Etienne Duguet,

Sébastien Lecommandoux and Christophe Schatz, “Control of the PEO Chain

Conformation on Nanoparticles by Adsorption of PEO-block-Poly(L-lysine) Copolymers

and Its Significance on Colloidal Stability and Protein Repellency”, Langmuir, 2011,

10.1021/la202990y

40. Kenneth J. Rosenberg, Tolga Goren, Rowena Crockett and Nicholas D. Spencer, “Load-

induced transitions in the lubricity of adsorbed poly(L-lysine)-g-dextran as a function of

polysaccharide chain density”, ACS Applied Materials and Interfaces, 2011,

10.1021/am200521m

41. Prathima C. Nalam, “Polymer brushes in aqueous solvent mixtures”, PhD Thesis No.

20211, ETH Zürich, 2012, 10.3929/ethz-a-007595210, page 87

Page 110: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

109

42. Prathima C. Nalam, Leonid Daikhin, Rosa M. Espinosa-Marzal, Jarred Clasohm,

Michael Urbakh and Nicholas D. Spencer, “Two-Fluid Model for the Interpretation of

Quartz Crystal Microbalance Response: Tuning Properties of Polymer Brushes with

Solvent Mixtures”, Journal of Physical Chemistry C, 2013, 10.1021/jp310811a

43. Laurent Feuz, “On the conformation of graft-copolymers with polyelectrolyte backbone

in solution and adsorbed on surfaces”, PhD Thesis No. 16644, ETH Zürich, 2006,

10.3929/ethz-a-005210741, page 236

44. Maximilian W. A. Skoda, Robert M. J. Jacobs, J. Willis and Frank Schreiber, “Hydration

of Oligo(ethylene glycol) Self-Assembled Monolayers Studied Using Polarization

Modulation Infrared Spectroscopy”, Langmuir, 2007, 10.1021/la0616653

Page 111: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

110

Chapter 7: Thesis Summary and Outlook

This project was motivated by the observation of ultralow friction and wear in biological

tribosystems, and in hydrophilic polymer brushes under aqueous conditions. The scope of the

thesis was to understand the impact of incorporating sugar-bearing chains into synthetic

brushes and thin films, thereby mimicking some of the mechanisms of natural lubrication.

In the first part of the project, surface-density gradients of dextran-bearing bottlebrush

copolymers were fabricated and the friction response measured using colloidal-probe atomic

force microscopy in aqueous solution. The results showed that the lubricity of dextran

brushes varies over several orders of magnitude and is dependent on both the chain density

and the applied load, with a characteristic transition between ultra-low- and high-friction

regimes manifesting above a certain applied load, which was also found to depend on chain

density. The explanation for this phenomenon was that inter-chain hydrogen bonding lent the

polysaccharide brush resistance against collapse, yet exacerbated friction forces upon

collapse and subsequent dragging on the counterface or interpenetration with a brush on the

opposing surface. This behavior is comparable to that of neutral brushes of sufficient length

and flexibility to entangle with the opposing brush. While the transition to high friction under

higher applied loads would seem detrimental to natural tribosystems or their future

biomimetic counterparts, the transition to the high friction regime occurs at much higher

pressures than are encountered in natural tribosystems, such as joints. Applications of related

materials could either avoid these high loads, or else use them to deliberately switch to high-

friction behavior at a certain applied load, which is tunable via the chain grafting density.

Upon discovery of this behavior, a new set of aqueous friction experiments on dextran

brushes was performed with kosmotropic additives in solution, in order to measure the effect

of a reduction in the strength of the inter-chain hydrogen bonding. The addition of moderate

amounts of potassium sulfate and trehalose in solution was found to shift the friction

transition to lower applied loads, while having less effect on the friction at low loads than a

corresponding reduction in brush density. This illustrated the relative contributions of inter-

chain hydrogen bonding and steric forces to the lubricity of the brush under different load

regimes. Since biological tribosystems of saccharide-bearing, brush-like molecules usually

operate in the presence of much higher salt concentrations, it is suggested that the presence of

other species that enhance the hydrogen bonding, or directly cross-link the chains, must serve

Page 112: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

111

to offset the failure of inter-chain hydrogen bonding under those conditions. Similar

techniques could be used to extend the lubricious regime of the dextran brush to higher loads,

or to selectively inhibit the resistance to collapse of the film in order to create a

heterogeneous film.

These findings led to a new line of investigation, to determine whether the thickness of the

layer and the brush conformation of the chains were influencing the hydrogen bonding, and

thereby the characteristic friction response of the system. To that end, a different set of

dextran-based brush-like films with higher chain densities, but less well-defined structure,

was synthesized and characterized. Dextran chains were attached at random intervals to the

surface by means of an intermediate UV-activated azide-bearing monolayer, and a moving

shutter was used to produce dextran surface density gradients. The results confirmed that the

brush conformation has a unique effect on friction, with the disordered chains displaying less

extreme friction behavior, exhibiting lower friction with increasing chain density but also

showing much weaker resistance to collapse on a per-monomer basis than the brush. For the

randomly attached system, significantly higher surface concentrations of dextran were

necessary to reach the same critical load (marking the transition from low to high friction

regimes) as in the original dextran brush, suggesting that the ordering of the polymer-brush

architecture significantly contributed the load-bearing properties of the dextran film.

Increasing the density of the disordered dextran film reduced the friction coefficient at high

loads, while the opposite was true for the polymer brush. Future brush architectures should

therefore aim to increase the film thickness and chain density without sacrificing the chain

ordering associated with the relatively uniform end-attached brushes – although this is a task

easier said than done.

Since the brush conformation was seen to play such an important role in friction compared to

less ordered hydrogen-bonding systems, the next goal was to understand the hydrated

conformation of the dextran chains. Computer simulations of coarse-grained dextran in

explicit water were used to study the equilibrium conformation of the brush, while optical

kinetics measurements were used to monitor the change in height and brush density during

the adsorption process. These experiments showed that the stiffness of dextran chains,

combined with the hydrogen bonding between chains, results in more extended brush

conformation than predicted by the scaling law for ideal brushes in good solvent. To compare

this behavior to the influence of hydration on chain conformation, the variation of PEG and

dextran brush structure under dry, humid and wet conditions was also studied via ATR-IR.

Page 113: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

112

Significant shifts in the C-O-C stretching bands were observed in PEG brushes with changing

humidity, indicative of conformational changes due to the presence of water. In contrast, no

corresponding shifts were observed in dextran, showing that the presence of water is

secondary to the inter-chain hydrogen bonding achieved by ordered chains in determining

brush structure, a factor which might explain the resistance to applied load exhibited by the

ordered brush in comparison to the disordered chains.

The findings of this work paint an interesting picture of polysaccharide brush lubrication. The

high availability, low toxicity and potential for extensive functionalization of polysaccharides

makes them exciting candidates for various industrial applications. The potential of achieving

ultra-low friction comparable to the best polymer brushes is present, but overloading the

polysaccharides, adding the wrong solutes or producing less regular brushes can quickly

change the situation for the worse. This implies that the use of comparable lubrication

mechanisms in nature must avoid or mitigate these potential pitfalls, for example by limiting

the operating pressure of tribosystems such as joints, exerting fine control over glycan

structure, stiffness and interactions, and/or physical or chemical crosslinking to strengthen

inter-chain interactions against the constant presence of high salt and other solute

concentrations. These insights will hopefully guide future studies of natural biolubrication, as

well as assist in the development of synthetic biomimetic lubricants.

Future work on heterogeneous architectures may allow for tailored films that exhibit the

appropriate mixture of properties, depending on the expected loading of the film. For

example, partial chemical crosslinking of dextran chains, using diisocyanates or another

active crosslinker, may lend increased load-bearing properties to the brush. It is possible that

even a small fraction of crosslinked chains might lead to a significant improvement in

resistance to spreading and interpenetration without reducing the hydration and lubricity of

the film, just as elastomers use a deliberately light degree of crosslinking to achieve their

unique properties. This method might also increase the brush resistance to collapse in the

presence of kosmotropes.

Brushes with heterogeneous chain compositions might also be synthesized using a grafting-

from method, in which parts of the chains are composed of hydrogen-bonding monomers

such as polyols, while other regions consist of a different monomer designed to bind water

without directly hydrogen bonding, such as PEG. Such a system might exhibit low friction

Page 114: In Copyright - Non-Commercial Use Permitted Rights ...8343/eth... · help with substrate coating and TInAS sensor ... When you blink, your upper eyelid slides along this ... which

113

even under high applied loads, and the “dual-action” mode of hydration and lubrication may

also reduce the friction in the high-load regime, for example in the presence of kosmotropes.

Ultimately, it is possible to imagine surfaces with heterogeneous composition, bearing

polysaccharide chains of different length, stiffness, degrees of crosslinking, and hydration.

They could be capable of either sliding past or trapping particles or biomolecules, perhaps

with some cyclodextrin-like side groups, or of achieving a tailored level of friction with

different counterfaces, sometimes entangling like Velcro and sometimes sliding like Crisco,

in response to the functionality of the opposing surface. While such surfaces are still far from

becoming a reality from a synthetic perspective, they have long been commonplace in the

eyes of Nature.